You are on page 1of 28

©2021 Society of Economic Geologists, Inc.

Economic Geology, v. XXX, no. XX, pp. X–X

Structural Controls on Alteration Stages at the


Chuquicamata Copper-Molybdenum Deposit, Northern Chile
Jorge Skarmeta†,*
Gerencia Corporativa de Exploraciones, Codelco-Chile, Huérfanos 1279 piso 8, Santiago, Chile

Abstract
All existing bench and tunnel vein and fault structural data with identified mineral infill, acquired in Chuquica-
mata, were georeferenced, digitized, and, according to their mineralogy, assigned to one or more of the major
alteration events developed between 35 and 31 Ma. Veins and faults were separated into two main stages: (1)
the late magmatic and potassic stage that comprises the background potassic and the propylitic alteration and
(2) the hydrothermal stage composed by early (intense potassic), main (principal and late sericite; hydrothermal
stages H1 through H4), and late (advanced argillic alteration) hydrothermal events. The spatial distribution of
the propylitic to late-hydrothermal events that plotted within the major fault framework indicate these had
either permeable or impermeable (±barrier) behavior through time. The area of the deposit was divided into
600 square grids measuring 100 × 100 m, and a stress orientation analysis was carried out for every propylitic
to late-hydrothermal alteration event. The analysis indicates that the local principal horizontal stress (sH) tra-
jectories are nonlinear and noncoaxial through the successive alteration events, differing from the previous and
following stages, and in the majority of cases do not coincide with the approximate east-northeast orientation
of the inferred tectonic far-field stress orientation. The differences between the stress trajectories, away from
the far-field stress orientation throughout the evolution of the system, are considered to be principally related
to the dynamic variations experienced by the stress components, such as thermal-magmatic stresses linked to
temperature fluctuations due to cooling or heating by progressive igneous/hydrothermal activity and/or elas-
tic, overburden-related stresses associated with reaccommodations developed during uplift and erosion. The
estimated stresses resulting after erosional unroofing and decreasing temperature indicate that the maximum
horizontal stress varied as the system evolved from the commonly accepted depth of emplacement of ~6 km.
During the late magmatic, background potassic, and intense potassic stages, the calculated differential stress
was contractional, decreasing to an isotropic state at the contraction-extension stress reversal that hosted the
main hydrothermal H1 through H3 events, to finally become extensional at the shallow late-hydrothermal
event. The most significant mineralization occurred at the time of stress reversal, coincidental with the seric-
ite and quartz-sericite events (H1-H4), associated with hydrothermal fluid accumulation, overpressuring, and
multiple-orientated hydraulic fracture development.
The Chuquicamata study suggests that the local stress control involved in the emplacement of porphyry
copper systems is fundamentally related to variable and progressive heat energy release, associated with igne-
ous and hydrothermal activity, and to the elastic stresses derived from uplift and unloading, rather than to a
constant far-field tectonic stress. The continuous local stress fluctuations led to bulk stress readjustments and
cyclical stress-fluid interactions for local fault reactivation, damage zone modification, brecciation, permeability
creation/destruction, and fluid focusing, as well as the discharge of hydrothermal fluids throughout the evolu-
tion of the system.

Introduction The Chuquicamata deposit has been exploited for over


The Chuquicamata Cu-Mo porphyry is among the largest 100 years. The pit is 6 km long, 4 km wide, and a little more
and richest deposits in the world (McMillan and Panteleyev, than 1 km deep. As the life of the pit is close to its end, pro-
1985; Cooke et al., 2005), with over 7.52 Gt of contained Cu duction will continue in an underground mine (currently
(produced plus resources) that corresponds to ~10% of the under construction) that will be developed with four levels
Chilean resources (Singer, 1995; Camus, 2003; Singer et al., of block caving that will extend 800 m below the current pit
2008). It is located within a highly endowed metallogenic floor. As of 2019 block caving will gradually replace existing
belt, which extends from southern Peru to central Chile along open-pit production, with a production rate of 366,000 tons
the western flank of the Andes Mountains (Maksaev, 1990; (t) of fine Cu per year, reaching full production plans after an
Camus, 2003; Sillitoe and Perello, 2005). This belt has an estimated nine years.
upper Eocene-Oligocene age (~31–45 Ma, bracketed within The engineering and planning studies for the underground
the Incaic tectonic event of the Andean orogeny, Noble Chuquicamata project required significant geologic infor-
et al., 1979; Cobbold et al., 2007) and includes Collahuasi, mation—in particular, 3-D structural models of the major
Quebrada Blanca, El Abra, Radomiro Tomic, Chuquicamata, internal structures. It also required databases with all exist-
Ministro Hales, and La Escondida giant deposits (Fig. 1). ing analogue information acquired from tunnels, bench faces,
and drill holes in a uniform digital and georeferenced format,
including the results of a set of deep drill holes especially
†E-mail:jskarmeta@gmail.com
undertaken for this purpose.
*Present address: Skarmeta Geological Services SpA, Julia Bernstein 720-N, A structural study to meet the project needs was designed
La Reina, Santiago 7850000, Chile. and implemented in 2013 (R. Cuadros, unpub. data, 2013). As

ISSN 0361-0128; doi:10.5382/econgeo.4769; 28 p. 1 Submitted: March 19, 2019 / Accepted: May 11, 2020

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
2 JORGE SKARMETA

show that body-derived stress components affected the total


70° PERÚ 69° stress budget. These include stresses developed during intru-
sion (Vigneresse et al., 1999; Tibaldi, 2015) together with the
18°

BO
thermal and erosional stresses (Jaeger, 1964; Cathles, 1977)

LIV
Arica developed while the deposit was being uplifted, eroded, and
Arica-Parinacota

IA
Region
telescoped into a condensed structural level (cf. Sillitoe, 2010).
CH

► Geology
EN
TR

Tarapaca Region Geodynamic context


20° Mesozoic-Cenozoic growth of the Andes involved varied
Iquique
tectonic regimes that reflect contrasting levels of mechani-
Queb. Blanca cal coupling along the convergent plate boundary (Mpodozis
Collahuasi
and Ramos, 1990; Mpodozis and Cornejo, 2012; Chen et al.,
2019). In the late-middle Eocene to early Miocene, the cen-
tral Andes (15°–25°S) experienced a phase of flat-slab sub-
El Abra
duction (Gutscher et al., 2000; Kay and Mpodozis, 2001) that
22° Tocopilla
CHUQUICAMATA boosted coupling and shortening along the plate margin, con-
Calama temporaneous with neutral or extensional conditions in the
southern Andes (De Celles et al., 2015; Horton, 2018). 
CHILE

The Domeyko Cordillera has developed as the result of a


Antofagasta Region
ATACAMA SALT LAKE phase of increased shortening during the late Cretaceous-
Antofagasta Cenozoic Andean orogeny, a consequence of the interac-
tion of the ENE-oriented convergence and the geometry
CHILE

24° Escondida
A of the plate margin in the central Andes. The Incaic event
E NTIN (45–30 Ma) is believed to represent the ultimate cause of the
ARG generation of several localized, relatively narrow orogenic
Metallogenic Belts belts, such as the Domeyko Cordillera and the Eastern Cor-
PERU

Miocene
dillera in Bolivia (Maksaev and Zentilli, 1999; Arriagada et al.,
2008). At the plate scale, the Incaic event does not correlate
Upper Eocene -
lower Oligocene
with any marked change in convergence vector but does cor-
26° Salvador relate with a marked slowing in convergence rate, which aver-
Paleocene -
Potrerillos aged between ~10 and ~5 cm/yr (Pardo-Casas and Molnar,
lower Eocene
Lower Cretaceous 1987; Somoza, 1998; Somoza and Ghidella, 2005; Sdrolias and
Caldera
Muller, 2006). The geodynamic relationship between slowing
Copiapo convergence, orogenic shortening, and increased east-west
0 50 100Km
compression along the South American margin (Silver et al.,
28° Atacama Region 1998; Sdrolaris and Muller, 2006) is unclear, although por-
70° 69° phyry ore formation seems to be intimately associated with
a high degree of intraplate coupling that can be explained by
Fig. 1. Location of Chuquicamata within the upper Eocene-lower Oligocene
(~40–30 Ma) metallogenetic belt. The map also shows belts of other ages that
orogenic shortening and slab shallowing (e.g., Mpodozis and
have been recognized in the area (see Camus, 2003). Inset shows location of Cornejo, 2012).
study area within northern Chile and South America. Causal relationships between magmatism, fertility, and
tectonics have been suggested for porphyry copper emplace-
ment (Kay and Mpodozis, 2001; Camus, 2003; Mpodozis and
a by-product of this operation-oriented study, the major faults Cornejo, 2012), and the fertility of the Incaic event appears to
were modeled and classified according to their degree of per- have progressively increased during its >10 m.y. life (Loucks,
meability and capacity to exert control on fluid flow. Based on 2014). Specific studies at El Abra (50 km north of Chuquica-
crosscutting relationships and mineral overprinting textures, mata; see Fig. 1; Rabbia et al., 2017) show that the magmatic
the vein populations and associated alterations were differ- history abruptly changes at ~38.7 Ma from normal (nona-
entiated in age, ordered into sequential stages, and plotted dakite) to adakite-like arc geochemistry (cf. Castillo, 2012),
within a spatial framework defined by the major faults. This with a significant depletion in rare earth element contents
study concluded that the major faults had controlled vein and (Gutscher, 2002; Rabbia et al., 2017). The evolution from non-
alteration distribution and that the local stress orientations, adakite to adakite-like magmatism coincides in time with the
determined based on vein orientation, varied significantly Incaic orogeny (Pilger, 1984; Pardo-Casas and Molnar, 1985;
throughout the period of alteration (~34–31 Ma) and do not Somoza, 1998; Somoza and Ghidella, 2012), which led to the
readily reconcile with the pattern expected for a constant and shortening and thickening of the arc’s crust, forcing lower
coaxial far-field stress at that time. In this paper it is proposed crustal magma chambers to evolve under higher pressure
that, given a likely constant tectonic (far-field) stress, the conditions, and to a slowdown of magma ascent, an increase
observed local variations in stress orientation and magnitude in magma residence time, and a change in geochemical

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 3

fractionation processes (Kay et al., 1999; Richards, 2003; Rab- equivalent to the Caracoles Group (Montaño, 1976), crops
bia et al., 2017). out and unconformably overlies the Paleozoic basement and
the Paleozoic-Triassic Collahuasi Group. Overlying the Juras-
District geologic setting sic rocks are Lower Cretaceous continental sandstones of the
The Chuquicamata district hosts the Chuquicamata, Radomiro Cerritos Bayos, the volcano-sedimentary rocks of the Upper
Tomic, and Ministro Hales porphyry copper deposits and Cretaceous Cerro Empexa, and the Eocene Icanche forma-
numerous smaller prospects (cf. Rivera et al., 2009, 2012). tions (Rivera et al., 2012). The Cerro Empexa Formation is
They are located in the late Eocene-early Oligocene (43–31 intruded by quartz diorite-monzonites of the Montecristo
Ma) porphyry copper belt (>1,000 km long) that defines the Intrusive Complex (U-Pb zircon age of 62.7 ± 0.5 Ma; Camp-
Domeyko Cordillera (Sillitoe and Perelló, 2005; see Fig. 1). bell et al., 2006).
The Domeyko fault system (Boric et al., 1990; Dilles et al., Eastern domain: A series of late Paleozoic granitoids have
1997) is a major feature that extends between 18°S and 31°S been grouped in the Cerros de Chuquicamata Plutonic Com-
and is 40 to 80 km wide. The Domeyko fault system is an plex (Fig. 2), which includes the Chuquicamata Hills diorite
orogen-parallel, discontinuous, overlapping, and segmented (273–267 Ma; Tomlinson et al., 2001), Mina Sur granodio-
array of bivergent dipping fault strands that juxtapose Paleo- rite, and the Mesa granite (~305 ± 4 Ma, K-Ar, Marinovic
zoic and Triassic basement blocks against younger folded and and Lahsen, 1984; U-Pb zircon age of 296.9 ± 2.1 Ma, Tom-
thrust Jurassic units and Cretaceous clastic and volcanic rocks linson and Blanco, 2008; Barra et al., 2013).  In the eastern
(McClay et al., 2002; Amilibia et al., 2008; Tomlinson and domain, east of Chuquicamata, the late Paleozoic and Triassic
Blanco, 2008). metasedimentary and volcanic rocks are intruded (Lindsay,
The West fissure as defined in Chuquicamata is an approx- 1998; Arnott, 2003; Tomlinson and Blanco, 2008; Rivera et al.,
imately north-south strand of the Domeyko fault system 2012) by the Triassic East granodiorite (229 ± 1 Ma, U-Pb in
that locally divides the district into eastern and western zircon, Tomlinson et al., 2001) and by the Elena granodiorite
geologic domains, from El Abra at the north end to Cerro (233–227 Ma U-Pb zircon ages, Tomlinson et al., 2001; J.M.
Limon Verde (~100 km) in the south. The nature and defor- Proffett and J.H. Dilles, unpub. report, 2007; Fig. 2).
mation history of this structure is disputed. Based on the Intrusive complexes: Eocene-Oligocene magmatism, occur-
geologic correlation between the Fortuna and El Abra intru- ring in both domains, is represented by the Los Picos, For-
sive complexes (Dilles et al., 1997, 2011), some authors tuna, and Chuquicamata intrusive complexes (Figs. 2, 3).
have proposed a multiphase tectonic history of Late Cre- The Los Picos Intrusive Complex, in the western domain, is
taceous to Oligocene margin-normal contraction combined the oldest unit with U-Pb zircon geochronology ages in the
with Eocene to mid-Oligocene margin-parallel dextral strike range of 43.1 to 41.8 Ma (Campbell et al., 2006; Barra et al.,
slip, which controlled porphyry emplacement and miner- 2013). Paleomagnetic results (Somoza et al., 2015) indicate
alization (Boric et al., 1990; Maksaev and Zentilli, 1999), that this plutonic complex underwent clockwise tectonic rota-
followed by ~37 km of Oligocene-Miocene postmineraliza- tion before it was intruded by the Fortuna Intrusive Com-
tion sinistral strike slip, as well as <1 km of late Miocene plex (~39 Ma). The Los Picos Intrusive Complex intrudes the
to Pliocene dextral strike slip (Taylor, 1935; Perry, 1952; Caracoles Group and Cerro Empexa and Icanche formations.
Reutter et al., 1991, 1996; Scheuber and Reutter, 1992; It is in turn cut by the Antena granodiorite (39.3 Ma, Dilles et
Lindsay et al., 1995; Tomlinson and Blanco, 1997; Tomlin- al., 1997; 38.8 Ma, Campbell et al., 2006; Fig. 2) and the volu-
son et al., 2001). Other authors (e.g., McClay et al., 2002; metrically dominant Fiesta hornblende granodiorite (U-Pb
Amilibia and Skarmeta, 2003; Amilibia et al., 2008) argue zircon ages of 37.6 Ma, Dilles et al., 1997; 39.4 and 38.2 Ma,
that the match between the Fortuna Complex and El Abra Campbell et al., 2006) that are intruded by the San Lorenzo
does not account for the ~2 km of vertical difference (and tonalite porphyries (38.2 Ma, Campbell et al., 2006). Based
consequent erosion) between both localities. Furthermore, on its hornblende barometry and mineral assemblage, Dilles
fission track determinations indicate that the Domeyko fault et al. (2011) suggest that the Fiesta unit solidified at ~700°C
system exhibits differential uplift with an alternate sense of at depths in the range of 4 to 7 km.
vergence along strike (McInnes et al., 1999; Olivares, 2001). In the Chuquicamata open pit the Chuquicamata Intrusive
Instead, they suggest the Cu-bearing porphyry intrusions Complex is fault juxtaposed against the Fortuna Intru-
focus along major, inverted, extensional faults in the outer- ­sive Complex, which has been considered to be the unaltered
arc extensional zones of their hanging-wall folds, developed and uplifted root zone of the Chuquicamata porphyry (Sillitoe,
at the climax of the Incaic contraction (Noble et al., 1979; 1973; Parada et al., 1987). However, Rivera et al. (2012), based
Cobbold et al., 2007). on Dilles et al. (1997, 2011), suggest the Fortuna Intrusive Com-
Western domain: Paleozoic schists of the Limón Verde plex does not correspond to the upfaulted roots of Chuquica-
metamorphic complex crop out to the south of the western mata, but rather to the roots of a different porphyry system that
domain (Tomlinson and Blanco, 2008; Fig. 2). A Paleozoic- was transported south by large-scale, postmineralization move-
early Mesozoic volcano-sedimentary sequence of sedimentary ment on the West fissure (Dilles et al., 1997; Tomlinson and
and dacitic to andesitic volcanic rocks and breccias (~303– Blanco, 1997; McInnes et al., 1999; McClay et al., 2002). Paleo-
232 Ma by U-Pb in zircon; Camus, 2003), correlated with the magnetic studies carried out in the Fortuna Intrusive Complex
Collahuasi Group, is exposed near the Toqui cluster (Tom- led Astudillo et al. (2008) to suggest that in the absence of solid-
linson et al., 2001; Munizaga et al., 2008; Rivera et al., 2012; state deformation, the apparent but irregularly oriented coun-
Barra et al., 2013; Fig. 2). In the southern part of the west- terclockwise rotations of small blocks (>100°) could be related
ern domain a marine sequence of Jurassic sedimentary rocks, to a West fissure damage zone (~200-m width).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
4 JORGE SKARMETA

E 500.000 E 510.000 E 520.000

Radomiro Tomic
mine

N 7.540.000 N 7.540.000

Chuquicamata
mine

Chuquicamata

N 7.530.000 N 7.530.000
Sur mine

Ministro Hales
mine
f issure

Toki cluster
West

N 7.520.000 N 7.520.000

Calama

0 2.5 5 Km

E 500.000 E 510.000 E 520.000


Cenozoic Mesozoic Paleozoic
Alluvial and colluvial unconsolidated deposits Montecristo Intrusive Complex: Collahuasi Group: andesites, dacites,
(Neogene) (lower Paleocene) sandstones (late Paleozoic -Triassic)
El Loa Group: limestones and siltstones Cerro Empexa Formation: Cerros de Chuquicamata Plutonic Complex:
(late Miocene - Pliocene) andesites, dacites, andesitic tuffs diorites and granodiorites (Permian)
Calama Formation: gravels (Upper Cretaceous)
Limon Verde Metamorphic Complex:
(Eocene - Oligocene) Tolar Formation: continental sandstones mica and amphibole schists (Paleozoic)
Chuquicamata Instrusive Complex (CIC): (Upper Cretaceous)
includes the East, West and Banco Cuesta de Montecristo sequence: Waste Dump
porphyries (35-34 Ma) andesites, dacites and continental
sedimentary rocks (Lower Cretaceous) City
Fortuna Intrusive Complex (FIC):
includes Tetera, San Lorenzo, Fiesta and Caracoles Group: marine limestones
Antena granodiorite porphyries (39-36 Ma) Unconformity
and sandstones (Jurassic)
Los Picos Intrusive Complex (LPIC): Fault, indicates sense of movement
Monzodiorites (44 - 41 Ma) East Granodiorite (Triassic) Reverse fault
Icanche Formation: andesites, dacites, and
breccias (early - middle-Eocene) Elena Granodiorite (Triassic) Inferred fault

Fig. 2. Chuquicamata district geologic setting (simplified from Barra et al., 2013). The West fissure divides the area into a
western and an eastern domain.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 5

Fortuna Intrusive Complex E4000


E3000

Pa
C2

niz
San Lorenzo N6000

W. F

o-
goza
porphyries

Po
Tetera Aplitic porphyry

issu

Zara

d
N

ero
Fiesta granodiorite N

re
B

s
E

a
S
B
E
N5000 N5000

re
W. Fissu

N4000 N4000
na

NN
ca
eri

500m
Am

Ar Chuquicamata
se Intrusive Complex
ni
co
Gravels
Brecciated
N3000 Qtz-Ser rock
Banco porphyry

West porphyry
East porphyry, dots
denote fine texture
Elena granodiorite

East granodiorite
lo
Faults and fault-veins
e zue E4000 Limestones
Denotes fault- vein zones P ort shales / sandstones
E3000
Fig. 3. Geologic map of Chuquicamata porphyry. The deposit is divided by a north-south postmineral structure (West fissure)
that juxtaposes the Fortuna Intrusive Complex and the Chuquicamata Intrusive Complex, which hosts the Chuquicamata
deposit. Rock types and major structures as exposed in the open pit in 1998 (simplified from Ossandon et al., 2001). True
north is 10° west of mine north. Faults in white, not mapped areas, were interpreted from rock mechanics data by Ossandon
et al. (2001). Abbreviations: EBN = Estanques Blancos North, EBS = Estanques Blancos South, Qtz = quartz, Ser = sericite.

Almost the entire Chuquicamata orebody is hosted by the 2006). The Chuquicamata Intrusive Complex intrudes the
Chuquicamata Intrusive Complex (East, West, and Banco Late Triassic Elena granodiorite (Figs. 2, 3; Lindsay, 1998;
porphyries; Fig. 3) in the eastern domain. Detailed descrip- Arnott, 2003; Zentilli et al., 2018).
tions of these units have been reported by Ossandon et al. The Fortuna Intrusive Complex in the western domain and
(2001) and Camus (2003), with U-Pb zircon ages that range the Chuquicamata Intrusive Complex in the eastern domain
from 34.5 to 33.1 Ma (Ballard et al., 2001; Campbell et al., have prominent foliations that support a syntectonic/plutonic

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
6 JORGE SKARMETA

emplacement (Reynolds et al., 1998; J. Dilles, unpub. report, a deep magma ± volatile zone at lithostatic, eventually supra-
2008). Somoza et al. (2015) established that the Fortuna Intru- lithostatic pressure, an overlying zone of transiently ascend-
sive Complex’s magnetic-tectonic foliations are parallel to the ing magmatic-hydrothermal fluids at temperatures ~400° to
magmatic fabrics and, therefore, were stress controlled during 650°C, and an upper brittle zone at temperatures <400°C,
emplacement and cooling. Ages of foliations show that the mag- characterized by a hydrostatically pressured mix of magmatic
matic and high-temperature hydrothermal system extended and nonmagmatic fluids (Tosdal and Dilles, 2020).
from ~36 to 33.5 Ma, while the sericite alteration and high- Chuquicamata exhibits a lateral distribution of alteration
sulfidation events have an age of ~32 Ma (Figs. 3, 4; Reynolds assemblages (Table 1). From the West fissure to the east it
et al., 1998; Ballard et al., 2001; J. Dilles, unpub. report, 2008). ranges from argillic spatially associated with enargite-sphaler-
The foliations are cut by hydrothermal tourmaline veins with a ite veins to a massive quartz-sericite zone along the deposit’s
gently plunging lineation (J. Dilles, unpub. report, 2008). western edge, transitioning eastward to a zone of discontinu-
Thermochronology data shows that as a consequence of ous remnants of the potassic alteration that mix with chlo-
uplift associated with the Incaic orogeny (Lamb and Davis, rite assemblages exposed along the eastern boundary of the
2003) there was rapid cooling between 35 and 30 Ma (Mak- deposit (Alvarez and Flores, 1985; Ossandon et al., 2001;
saev, 1990; McInnes et al., 1999). These calculations yield Faunes et al., 2005; Fig. 4). These overlapping relationships
exhumation rates in the order of 200 m/m.y. (Maksaev and indicate superimposition of successive incremental pulses of
Zentilli, 1999), which decrease to ~50 m/m.y. after 30 Ma. By fluid introduction that characterize the different hydrother-
12 Ma, the Chuquicamata Intrusive Complex and contained mal stages, simultaneous with overprinting of earlier material,
mineralized porphyry systems were exposed to surface oxidiz- uplift, unroofing, erosion (cf. England and Molnar, 1990), and
ing conditions (Maksaev and Zentilli, 1999). These conditions increase of the sulfidation stage, interpreted to be the result of
persisted until precipitation drastically changed from >200 a shallowing emplacement depth, cooling, and embrittlement
to <20 mm/yr (between 19 and 13 Ma; Rech et al., 2006), of the host rock (Figs. 4–6). 
broadly coinciding with ages determined from supergene The hypogene alteration has been separated into two major
alunites (19–15 Ma; Sillitoe and McKee, 1996; Sillitoe et al., stages, which decrease in age and correspond to the late
1996; Pinget et al., 2015) and U-Pb crystallization ages from magmatic-potassic and the hydrothermal events, both with
pseudomalachite bands that yield 18 ± 1.0 Ma (Kahou et al., internal subdivisions and characteristics (cf. Sillitoe, 2010;
2020, Fig. 4). This uplift generated Miocene synorogenic Fig. 4; Table 1). The potassic alteration has been divided in
gravels (i.e., El Loa Group; Blanco, 2008) into which copper the early background potassic and propylitic (Fig. 4; Table 1).
solutions sourced from Chuquicamata and Radomiro Tomic Background potassic alteration developed between 34 and
deposited as exotic mineralization (Jarrel, 1944; Munchmeyer, 33.4 Ma, is younger than the East porphyry host rock, and is
1996; Cuadra et al., 1997; Ossandon and Zentilli, 1997). Post- younger than the West and Banco porphyries that intrude the
mineral Miocene continental clastic sediments and younger East porphyry (Ossandón et al., 2001; Arnott, 2003; Arnott
ash and ignimbrite intercalations are widely distributed across and Zentilli, 2003; Faunes et al., 2005; Rivera et al., 2012; Fig.
the district (May et al., 2005; Fig. 2). 4), and propylitic alteration is transitional and distal to the
potassic-altered rocks (Table 1). The hydrothermal activity
Hypogene Alteration-Mineralization has been divided into early, main, and late stages (i.e., Ossan-
The Chuquicamata alteration-mineralization pattern com- don et al., 2001; Camus, 2003; Faunes et al., 2005; Figs. 4,
prises, from bottom to top, potassic, chlorite-sericite, seric- 6). The early hydrothermal corresponds to the intense potas-
ite, and advanced argillic zones (Meyer and Hemley, 1967; sic sericite (±quartz) or phyllic alteration characterized by
Lowell and Guilbert, 1970; Seedorff et al., 2005), with two the partial or complete destruction of the original rock tex-
overlapping and transitional stages between them (Faunes et tures (Fig. 4; Table 1). The main hydrothermal stage has been
al., 2005; H1 and H3 in Fig. 4 and Table 1). divided into a principal (H2) and a late quartz-sericite (H4)
The overall magmatic-hydrothermal activity lasted approxi- stage and their transitional stages (H1 and H3; Fig. 4; Table
mately ~2 m.y. (Fig. 4). Whether the Chuquicamata hydro- 1). The late-hydrothermal stage represents the last superim-
thermal mineralization developed in a continuous or in a posed advance argillic alteration stage (Table 1) composed of
punctuated way is yet to be determined and beyond the scope massive high-sulfidation veins. 
of this study. However, the hydrothermal alteration assem- Detailed descriptions of the different alteration stages have
blages (Fig. 4; Table 1) indicate it developed in superimposed been documented by Ossandon et al. (2001), Faunes et al.
increments, in a way similar to the protracted, but episodic, (2005), Cordova et al. (2010), and Rivera et al. (2009). Their
magmatic histories described in many Andean, North Ameri- main characteristics are summarized in Table 1 and Figure 4.
can, and Philippine porphyry Cu districts (e.g., Cornejo et al.,
1997; Richards et al., 2001; Harris et al., 2004; Padilla-Garza States of Stress Driving Vein-Mineralization:
et al., 2004; Barra et al., 2005; Deckart et al., 2005; Sillitoe A Preamble
and Mortensen, 2010; Simmons et al., 2013; Braxton et al., Veins and stockworks host 80 to 90% of the mineraliza-
2018). Moreover, the alteration patterns of Chuquicamata tion and alteration present in Chuquicamata (Perry, 1952;
(e.g., Camus, 2003) are similar to those classically described Ambrus, 1979; Soto, 1979; Lindsay, 1998). Stockwork and
in other porphyry systems (Meyer and Hemley, 1967; Lowell sheeted extensional veins are spatially distributed near low-
and Guilbert, 1970; Seedorff et al., 2005; Sillitoe, 2010) and displacement fault fractures (Fig. 3) and share a common
have been grossly correlated with fluid and thermal regimes hydrothermal mineral infill that indicates synchronous struc-
(Monecke et al., 2018; Tosdal and Dilles, 2020). These include tural development at the different alteration stages (Fig. 4).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 7

Cu % / Total sulf. content


Sulfidation and Qtz
Mineral

Temp °C range
stages/events
Rock Textures

fault behavior
Acidity state
Vein classificaon associaon,

Impervious
alteration

Colo r Codes

Depth Km
Major
Alteraon
Pyrite
Stage
Age and mineral relave
(Ma) assemblage abundance
and type

15
Supergenic Processes
19
Massive E veins, 0
High (HS > 5%) and Qtz

31,1

and Estanques
Americana
Advanced High intensity Qtz with illite-sericite

250 - 180

N and S.
Blancos
Py 80%;
argillic LH -alunite halos, Py
Late

< 0,2
and Sb, As sulfides Cu sul < 20%
alteraon -En in stockworks.
Py - En Tenn Tetr.
High

1
Pervasive QSP altera­on

Blancos N.
Americana
Late quartz D veins, with Py-Cpy

Estanques
sericite, H4 (late phyllic) 80%. Total

270 - 220
0,1 / 2%
in sutures and Py >> Cu sulf

and
Phyllic replacement of feldspars sericic halos
Total destruction

(QSP) and bio­te by sericite


Hydrothermal

Argillic Possibly
Transi onal H3 Qtz - Py Cpy Sph. Americana
(<20%) 2 and Zaragoza
Intermediate (1 - 3%)
Intermediate
Main

Coarse early green - C veins with green


principal sericite

550 - 500
> 0,5 / > 1%
Cpy > 80%;
Intermediate

gray sericite and Cpy Zaragoza,


H2 disseminated and in vein disseminated Cpy Py < 20% Panizo
halos; veins with and/or K Feldspar Bn absent and C2
central suture; peripheral
halos; central
to the Intense Potassic.
suture scarce.
Qtz-Mo Phyllic alteraon 3 Zaragoza,

~ 550
High Mo (01-02%)
32,9 Transional H1 B (“blue”) veins

< 0.6
Superimposed on EH Low Py and Panizo
Moderated to intense
Partial destr.

Irregular
Qtz appears

Intense replacement of feldspars and A veins 30%; distribu on Estanques

550 - 400
Early

1 / > 1%
Low

potassic EH bio­te by secondary K feldspar Blancos N.


Qtz - Mo
and green sericite. KSil 30%
Cpy, Bn Dg Cv Cpy
in cataclas­c banded veins Py absent.
Propylitic

0,2 - 0,5 /< 0.6%


Late magmac and potassic

Selecve ( 50%) alteraon 450 - 350


PR of mafic minerals; primary Clorite (secondary Dissemina on
Low (LS< 1%);Qtz absent

biote) and specular None


magnete altered to clorite Py > 50% > Cpy
hemate veins
Preserved

4
Disseminated biote; Mainly
and
Background

0,2 - 0,5 / < 1%

33,4 100% hornblende disseminaon


potassic

650 - 550

5 ?
altered to biote;
Absent

BK Early quartz - Cpy + Bn


magnete present; sulfide free
>50%;Cpy>Bn
6
33,5
West and Banco Porphyries
34,5
East Porphyry
Magmatic

36,2
Fortuna Complex alteraon
38,0
Precursor Fortuna Complex
39,5
Fig. 4. Time-alteration chart showing the eight major alteration stages (background potassic to late hydrothermal), alteration
descriptions, vein distribution and classification, mineral association, and estimated temperature ranges for each mineral
association (see text for details). Abbreviations: BK = background potassic, Bn = bornite, Cpy = chalcopyrite, Cv = covellite,
Dg = digenite, EH = early hydrothermal, En = enargite, H1-H4 = main hydrothermal stages, HS = high-sulfidation, KSil =
local name for moderate to intense replacement of feldspar and biotite by secondary K-feldspar (from Faunes et al., 2005),
LH = late hydrothermal, LS = low-sulfidation, PR = propylitic, Py = pyrite, QSP = quartz-sericite-phyllic, Qtz = quartz, Sph
= sphalerite, Sulf = sulfidation, SVG = gray-green sericite (from the Spanish sericita verde gris), Tenn = tennantite, Tetr. =
tetrahedrite.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
8 JORGE SKARMETA

Table 1. Summary of Hypogene Alteration Characteristics Listed from Top to Bottom


(based on Ossandon et al., 2001; Faunes et al., 2005; Rivera et al., 2009; Cordova et al., 2010)

Stage Events Code Distribution Mineral and textural characteristics

Hydrothermal
Late Advanced argillic alteration LH Exposed between the West fissure and the The alteration assemblage is defined by the
  Americana fault; extends to the north up to   presence of kaolinite, montmorillonite, and  
  the Estanques Blancos Norte fault   sericite with traces of pyrophyllite and/or  
  (Figs. 3, 4, 6)   dickite locally with hypogene alunite  
  (Ossandon et al., 2001); clay species replace  
  feldspar and mica

Main Late quartz sericite, phyllic H4 Spatially distributed to the west of EH (Fig. 6) Aggregates of sericite, quartz, and pyrite (QSP)
  that obliterate the original textures and bleach  
  the rocks  




Transitional H3 Mainly exposed at the NE sector of the Quartz, sericite, pyrite, chlorite, and clay
  deposit (Fig. 6)   aggregates (kaolinite, illite-smectite) completely  
  obliterate lithological textures and replace  
  feldspars and biotite (Faunes et al., 2005)  

Intermediate principal sericite H2 Peripheral to EH (Fig. 6) Sericite and biotite in alteration halos
  (Faunes et al., 2005; Ossandon et al., 2001)  








Transitional H1 Present in the center of the deposit, The phyllic alteration overprints the BK and PR
  superimposed on the intense potassic event;   alteration; lithological textures are preserved,  
  coexists with the late hydrothermal stages   and plagioclase and K-feldspar are partially  
  (Figs. 4–6)   replaced; the bulk of the phyllic alteration  
  occurs in vein and veinlet selvages between  
  which the potassic alteration is preserved  

Early Intense potassic EH This alteration event is distributed along the Partial or total destruction of rock textures,
  central part of the deposit (Fig 6)   giving rise to a fine-grained gray rock, referred  
  to as Ksil; the Ksil comprises moderate to  
  intense replacement of feldspar and biotite by  
  secondary K-feldspar, in some instances  
  accompanied by secondary albite and quartz  
  (Faunes et al., 2005)  

Late magmatic and potassic


Propylitic PR Transitional and distal to the potassic alteration Selective alteration (~50%) of magmatic
  and peripheral to the background potassic,   amphiboles, biotite and magnetite to chlorite.  
  beyond the economic limits of the deposit   Distal epidote occurs at the deposit margins,  
  (Fig. 4); outer limit is defined by the   and the rock textures are well preserved  
  occurrence of unaltered hornblende  

Background potassic BK The patchy distribution suggests it corresponds Primary hornblende is altered to secondary
  to relicts of a widespread distributed alteration   biotite; feldspars are generally unaltered, and  
  that was subsequently obliterated (Ossandon   magnetite is destroyed or altered to fine-grained  
  et al., 2001, Lindsay, 1998; Figs. 4, 6); outer   hematite  
  limit is defined as where hornblende has not  
  been biotitized (cf. Seedorff et al., 2005)  

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 9

Table 1. (Cont.)

Stage Ore mineralogy and veins Equilibrium temperature estimations Age

Hydrothermal
Late Massive high-sulfidation E-type veins
  (Masterman et al., 2005; Maydagán et al.,
  2015), with pyrite-enargite-chalcopyrite-
  bornite-digenite and illite-sericite in
  the halos

Main D-type veins with pyrite (~50%) > Stable isotope studies indicate that the fluids Based on 40Ar/39Ar and Re-Os ages this event
  chalcopyrite; chalcopyrite is mostly in the   associated with the sericite alteration were   was assigned to the 32–31 Ma age range
  central veinlet and less abundant in the   a magmatic-meteoric mix (Arnott, 2003);   (Reynolds et al., 1998; Arnott, 2003; Barra
  sericite-quartz-chalcopyrite halos (Faunes   sulphur isotope (Lewis, 1996) and fluid   et al., 2013), 1 m.y. younger than the B-type
  et al., 2005; Sillitoe, 2010; Fig. 5)   inclusion studies (S. Collao, unpub. report,   veins, 2.5–3 m.y. younger than the BK event,
  1987; Vega, 1989) suggest temperatures in   and 3.5 m.y. younger than the crystallization
  the range of 210°–470°C   age of the East porphyry (Figs. 3–5)
Late pervasive quartz-sericite argillic Clay minerals present (rare dickite) and
  alteration characterized by abundant pyrite   absence of sericite and pyrophyllite restricts
  and Cu sulfides that superpose on, and   argillic alteration to temperatures below
  reutilize, earlier high-sulfidation pyrite veins   200°C (Meyer and Hemley, 1967; Arancibia
  and Clark, 1996; Sillitoe, 1993, 2010)
Irregular C-type veins and a pervasive Crosscutting relationships between quartz-
  dissemination of fine chalcopyrite with   molybdenite B-type veins and high-pyrite
  coarse-grained gray quartz sericite halos   D-type veins (Fig. 5) associated with the late
  (Fig. 6); veins contain varying proportions of   quartz-sericite events demonstrate their
  quartz, Cu-Fe sulfides, enargite, tennantite,   relative earlier age (Faunes et al., 2005)
  and sphalerite; where gray-green sericite
  content is >10% the total sulfide is >1%, and
  Cu grades are well above 1%, making this
  the principal hypogene mineralization event
  at Chuquicamata (Fig. 4)
Banded quartz blue veins (B-type, <1 cm) By analogy with fluid inclusion studies in Re-Os ages on molybdenite from B-type
  that crosscut early-stage quartz and quartz-   B-type veins from Butte, Montana (Rusk   veins range between 33 and 32 Ma
  K-feldspar A-type veins; they are in turn   et al., 2004, 2008) the temperature range   (Ossandon et al., 2001; Zentilli et al., 2015)
  cut by larger banded quartz B-type veins   varies from 500°–600°C, in agreement with
  (0.5–1 m) with abundant molybdenite   the ~550°C estimated by Seedorff et al.
  (Lindsay et al., 1995; Fig. 5); corresponds   (2005)
  to the principal Mo deposition at
  Chuquicamata (Arnott, 2003; Fig. 4)

Early Early-stage quartz-K-feldspar A-type veins


  and pervasive quartz-K-feldspar alteration
  (Faunes et al., 2005; Fig. 4); primary por-
  phyry feldspars and biotite are completely
  replaced by K-feldspar and an aggregate of
  quartz and sericite associated with very fine
  grained disseminations of bornite ± digenite
  ± covellite, or chalcopyrite ± covellite +
  bornite ± digenite (Faunes et al., 2005)

Late magmatic and potassic


Primarily composed of chlorite veins The formation temperature decreases from
  (Ossandon et al., 2001; Faunes et al., 2005)   the potassic alteration down to ranges of
  and fracture surfaces with chlorite-magnetite-   350°–450°C (Titley, 1993; Reed, 1999,
  sericite; copper is disseminated, with pyrite >   see Fig. 4)
  chalcopyrite, and the total sulfide content is
  less than 0.6% by volume
Biotite may be finely disseminated or Sulfide phase equilibria (Lewis, 1996) and O Age determinations range between 34.0 and
  concentrated as irregular (<0.5 mm)   stable isotope analyses allowed Arnott and   33.4 Ma (Ar/Ar in biotite and K-feldspar;
  discontinuous veinlets (EB-type veins of   Zentilli (2003) to suggest that the potassic   Reynolds et al., 1998; Arnott, 2003);
  Gustafson and Quiroga, 1995), or as halos   alteration was in equilibrium with magmatic   Ossandon et al. (2001) and Lindsay (1998)
  along K-feldspar ± quartz (1 cm) veinlets   fluids at ~535°C, close to the temperature   conclude it is related to all three porphyries
  (M-type veins of Arancibia and Clark, 1996);   range (550°–650°C) commonly assigned to   (Fig. 3), although Arnott and Zentilli (2003)
  pyrite is usually absent or rare (Faunes et al.,   this mineral assemblage (Beane, 1974;   considered it is only related to the East
  2005; Cordova et al., 2010; Rivera et al., 2012)   Titley, 1993; Seedorf et al., 2005)   porphyry

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
10 JORGE SKARMETA

The stress states associated with the vein development varied accommodated by oblique low-displacement shear fractures
throughout the alteration cycle, as shown by the crosscutting (e.g., Hill-type structures; Hill, 1977; Sibson, 2001), all with a
chronology of the veinlet sequence (Figs. 5, 6). Vein orien- common hydrothermal mineral infill.
tations show high variability between the different alteration Mineralized veins and primary fracture permeability mesh
stages. These variations imply that the stresses at any one point development (ground preparation; Sibson, 2000, 2001) in the
in time of Chuquicamata development are the product of a overlying wall rock can be related to the mechanics of por-
complex combination resulting from the addition and subtrac- phyry intrusion and pressure-driven roof extension (Gruen et
tion of local stresses that encompassed its geologic evolution. al., 2010; Tosdal and Richards, 2001; Tosdal and Dilles, 2020).
Magma intrusion may enhance cracking and rock rupturing of
Extensional veins and hydraulic fracture development the cupola, overpressured hydrothermal and magmatic fluid
Sheeted extension veins that developed by overpressurized circulation (Gudmundsson, 2011), heat and mass exchange
fluids in brittle rock during hydrothermal alteration (propy- with the host rock, and rheological modifications that result
litic to late hydrothermal; Figs. 4, 6; Table 1) are dependent from heat and magma pressure increase (Gruen et al., 2010;
upon the relationship of differential stress (s1 – s3) to rock ten- Tosdal and Dilles, 2020).
sile strength, To (effectively, a measure of rock competence;
Secor, 1965; Phillips, 1972; Engelder, 1987, 1994), which in Reference and local stress fields
volcanic and crystalline rocks is in the range of 10 to 30 MPa Reference states of stresses describe the stress changes through-
(Birch, 1966; Lockner, 1995). out the crust, assuming it is not affected by tectonic forces, and
Within intact rock, the formation of extensional vein arrays to establish the tectonic stress it is necessary to determine the
occurs when (s1 – s3) < 4To, constraining the magnitude of extent of deviations from these reference models (McGarr and
differential stress during their formation (Etheridge, 1983). Gay, 1978; Gough and Gough, 1987; Fossen, 2016).
Extension vein formation is favored by high tensile strength Elastic state of stress: The elastic (or uniaxial) reference
and low differential stress (Secor, 1965; Sibson, 1985, 2001). state of stress is based on the boundary condition that no elon-
In contrast, the development of shear fractures containing the gation occurs in the horizontal directions, strain only occurs in
s2 axis is more likely to develop in low-tensile-strength rock the vertical direction, and stresses comply with this condition.
or where the differential stress is higher than 4To. Hence, During burial, uplift, and erosion, in contraction (s1 = sH) or
the differential stress and the degree of overpressuring transpression (s2 = sV), the principal horizontal stresses (sH =
becomes inversely interdependent, as high fluid overpres- s1 ≥ sh = s2 or s3) are equal and increase as a function of the
sures are incompatible with high differential stress and vice vertical stress (sV) and can be approximated by the following
versa (Etheridge, 1983). Mineralized veins (Fig. 4; Table 1) equation (Jaeger and Cook, 1979):
develop at extensional and extensional-shear stress levels,
                 
typically of small displacement and high dilation (Cox, 1999, s         v
H = sh = —— ⋅ sV, (1)
2010, 2020). The extension-shear or hybrid type of fractures                           1–v
develops in low-differential stress regimes, within the 4To ≤ where v is the Poisson’s ratio (~0.25), and, accordingly, the
(s1 – s3) < 5.6To range (Secor, 1965; Sibson, 2017). This occurs horizontal stresses will become about a third of sV (McGarr,
either as extension veins with oblique-to-wall fibers, in which 1988; McGarr and Gay, 1978). Nevertheless, data compiled
shearing is accommodated inside the vein itself, or by pure from the upper 2.5 km of the crust show that horizontal stress
extension veins with normal-to-the wall fibers that become becomes systematically larger than the vertical (Engelder,
1993; Sheorey, 1994; Hergert and Heidbach, 2011).
BK-PR; M; EB Regional (tectonic) stress fields: Tectonic stresses (st) cor-
respond to the stresses that cause a deviation from the ref-
H4; D erence state of stress, commonly related to plate boundary
LH; E 31-32 Ma tractions and plate tectonic movements. Although it is difficult
Quartz-sericite to estimate, the absolute magnitude of the tectonic stresses is
halo equivalent to the total maximum horizontal stress (sH) minus
all local stresses.
Current stress fields determined by in situ measurements
K-feldspar H1; B and active fault solutions show regional patterns with homo-
halo 32-33 Ma geneous and uniform orientations (Zoback et al., 1989; Zoback
and Magee, 1991; Zoback, 2010). It has been shown (Jaeger
and Cook, 1979) that the upper crust’s principal horizontal
stresses have an upper limit given by the preexisting critically
EH; A oriented faults that are at the edge of failure (McGarr, 2014;
Sibson, 2017) and by the crust’s frictional strength:
H2; C
          s 1   sH
Fig. 5. Schematic crosscutting chronology of veinlet sequences at Chuqui-
camata. Veinlet nomenclature follows Gustafson and Hunt (1975; A, B, and
          s
–— [ ]
= —– = (m2 + 1)1/2 + m 2 ≤ 3
3   sV
or
D types), Arancibia and Clark (1996; M and EB type), Dilles and Enaudi, sH ≤ 3 sV (2)
(1992; C type), and Masterman et al. (2005) and Maydagán et al. (2015; E
type). Association with alteration stages are color coded as in Figures 4 and 6.
for m ~0.55, (where m is the dynamic coefficient of friction).
Modified from Sillitoe (2010). Present-day inversion of earthquake focal mechanisms yields

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 11

Fig. 6. (a-h) Plan view of the distribution of alteration stages (propylitic-late hydrothermal) plotted over the very important fault (VIF) skeleton. Alteration stages abbrevi-
ated and color coded as in Figure 4.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
12 JORGE SKARMETA

a sH/sV ratio that varies from 1.5 to 2.5 with consistent orien- consequence of uplift and cooling, the thermal effect will be
tation (cf. Cui et al., 2016; Soh et al., 2018). essentially extensional.
Magmatic stresses: The propagation and plumbing of mag- Stress variations during uplift and erosion: By consideration
mas and the shape of the intrusions are essentially controlled of the thermal and elastic effects generated during exhuma-
by buoyancy forces, crustal heterogeneities, and the orienta- tion (e.g., England and Molnar, 1990; Wilkinson and Kesler,
tion of the extant far-field stress (Gudmundsson, 2011). Buoy- 2007), the resulting stress history can be estimated (eq. 1, 3).
ancy is the result of the density contrasts between the magma From a starting point at ~6 km (Seedorff et al., 2008; Proffett,
and hydrothermal fluids with respect to the country rocks. 2009; Sillitoe, 2010), the horizontal stress at different depths
If melt segregates to form veins, the maximum buoyancy will correspond to
stress may evolve in two ways. In dike-controlled geometries,
         1–2v       a⋅E
sH = sh = ——– ⋅ sV + —— ⋅ DT
a series of breccias and meshes of interconnected, steeply           1–v        1–v
dipping veins that crosscut across the wall rock will develop.
            Dz
Conversely, diapir-type intrusions are likely to rise in shorten- [
= —— ⋅ (1 – 2v)r ⋅ g + a ⋅ E ⋅ DT ,
            1–v ] (4)
ing environments, generating disconnected veins (Weinberg,
1996). The shapes of the Chuquicamata porphyries are dif- where Dz is the change in depth as a result of uplift and ero-
ficult to establish; however, in plan view they vary from dike- sion of the mineral system, r is the density of the host rock, g
like in the West and Banco to a diapir-type represented by the is the acceleration due to gravity, and ± DT (heating or cool-
East porphyry (Fig. 3). Calculating the real contribution of ing) is the temperature variation due to intrusion.
magma stresses is limited by the lack of host rocks preserved
on top of the porphyries. Structural Studies and Structure Classification
Studies from eroded igneous deep crustal volcanic edifices
(Tibaldi, 2015) show that the path of magma transport is gov- Methodology and databases
erned by the existing ambient stresses and that dikes propa- All geologic and structural data contained in hard copy maps
gate through fractures (Shaw, 1980; Spera, 1980; Petford, and mapping sheets available in 2013 from over 220,000 m of
1996; Weinberg, 1996). As in porphyry systems, the ambient bench face mapping and all underground tunnel maps were
stress fields get modified during intrusion and consequent compiled, digitized, and georeferenced in AutoCAD. All digi-
uplift (Tibaldi, 2015; van Wyk de Vriers and van Wyk de Vri- tized structures and their geologic and structural descriptions
ers, 2018). were stored in a standardized Excel database, which allows
Thermal effect on horizontal stress: The effect of temper- the use of specific software packages to interrogate the data
ature variations on horizontal stress by uplift, cooling, local (e.g., Dips, Leapfrog, or Vulcan).
magma intrusion, and/or hydrothermal heating may tran- When the information extracted from the mapping sheets
siently distort the reference state of stress and significantly allowed the classification of the vein mineralogical infill
modify the differential and deviatoric stresses. Intrusion (~60% of the data), the structural features were color coded
causes thermal stresses at different temperature and/or depth according to the different hypogene alteration stages (Fig.
(Rikitake, 1959; Jaeger, 1964; Cathles, 1977), where the over- 4). All data entries specify the coordinates, source, spatial
all temperature variation, after multiple heating/cooling epi- attitude of faults and veins, surface attributes and morphol-
sodes, is ± DT, and the corresponding thermal stress can be ogy, and alteration-mineral association parameters. A total of
obtained from 506 hard copy drill hole folders that contained core descrip-
tions and photographs were revised for the identification and
                       a⋅E
s = —— ⋅ DT, (3) selection of their intercepts with major very important faults,
                    T   1–v
from which 85 drill holes were selected for further structural
where E is Young’s modulus, a is the linear thermal expan- relogging. Based on the information from the relogged struc-
sion coefficient, DT is the temperature change, and v is Pois- tural intercepts, 151 sections were drawn across six major
son’s ratio. The depth-to-temperature ranges were assigned faults (very important faults; see below) for wire frame and
from the alteration mineral equilibrium assemblages (see solid constructions and 3-D modeling (Table 2) and classified
Hypogene Alteration; Fig. 4; Table 1) and correlation with in terms of their potential to influence underground mine
data from similar systems (e.g., Titley, 1993; Sillitoe and development.
Hedenquist, 2003; Rusk et al., 2004; Heinrich, 2005; Master- The principal alteration stages recognized throughout the
man et al., 2005; Seedorff et al., 2005; Richards, 2011). As a mine (Lindsay, 1998; Ossandon et al., 2001; Faunes et al.,
Table 2. Summary of Dimensions, Orientation, and Sections Used for Construction of 3-D Fault Zone Models

Fault zone Sections N coordinates Extension (m) Trend Dip Thickness (m) Depth (km)

Americana 56 2250–5000 3,000 16° E 80º–85º W 35 1.6


NNW 15 3700–4400 700 15° W 85º–88º E 15 1.7
EBN 15 5050–5700 1,150 55° E 70º–75° SE 28 1.7
EBS 15 4200–4980 1,000 45° E 75º–85° W 18 1.6–1.7
Zaragosa 23 5100–6200 1,100 15° E 85º W 28 1.7
C2 24 4350–5500 800 18° W 85º–88º W 21 1.7

EBN and EBS are Estanques Blancos North and South; see Figures 3 and 10

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 13

2005; Cordova et al., 2010; Rivera et al., 2012; Table 1) were unpub. report, 2013) are those that extend in pit exposures
arranged in a relative chronology on the basis of their cross- in excess of three benches and exceed 0.5 m in width of the
cutting vein relationships, characteristic infill, and absolute complete fault zone (core + damage zone, see below). Fault
age, where available (Reynolds et al., 1998; Barra et al., 2013; traces represent smaller structures, almost always spatially
Figs. 5, 6). associated with very important faults, that extend more than
The vein orientation and density analyses for all alteration one bench with have widths of less than 0.5 m.
stages were assessed by subdivision of the whole mine area into Very important fault systems develop heterogeneous fault
600 square (100 × 100 m) grids, each with an alphanumeric cores, composed of slip surfaces, breccias, and extension
nomenclature (A1-40 to O1-40). With the color-coded vein veins, and have a broader volume of distributed deformation
orientation data from each grid, a total of 548 rose diagrams (damage zone; cf. Faulkner et al., 2010). The most signifi-
were generated, per grid and per alteration event (propylitic cant very important fault structures (Tables 2, 3) were clas-
to late hydrothermal; Fig. 4; Table 1). In each rose diagram sified following a modification of Caine et al. (1996), which
the orientation of data was subject to the Terzaghi geometrical compares the percentage of the total width of the fault zone
correction (Terzaghi, 1965), a technique that modifies linear (fault core materials) to the percentage of subsidiary damage
and planar fracture samples with respect to directional bias zone structures. The fault core and damage zone are distinct
in order to establish a more precise orientation of the cor- structural units that reflect material properties, hydrological
responding structure with respect to the rock slope direction behavior, and deformation conditions of the fault zone.
(see example in Fig. 7).
Stress trajectories throughout alteration
Structural classification and very important Based on the corrected rose diagrams (Fig. 7), the principal
fault characterization horizontal stress orientations were determined for each alter-
The structural ore control in Chuquicamata is mostly in the ation stage in all of the 600 grid cells described above. The
form of fault-vein hybrid-type structures, where major ore- main principal paleostress orientation was inferred as paral-
hosting faults are typically low-displacement features (e.g., lel to the early extension fractures and/or as the bisecter of
Newhouse, 1940; McKinstry, 1948). They are dominated by the dihedral angle (±22°–32° centered on the s1 direction;
dilatational > wear-damage structures, genetically and spa- Angelier, 1994; Sibson, 2000; Fig. 8) of the conjugate shear-
tially associated with extensional veins and veinlet-hosted extension fractures (hybrid) that succeed and overprint the
stockworks that host 80 to 90% of the mineralization and extension veins, as fluid pressure drops and the stress differ-
alteration (Ambrus, 1979; Soto, 1979; Lindsay, 1998). These ence increases (Hubbert, 1951). Hence, small angles between
hybrid structures have traditionally been mapped and referred fracture sets indicate elevated fluid pressures (Secor, 1965;
to as very important faults and fault traces (Tables 2, 3). Very Engelder, 1987).
important faults (I. Collado, unpub. report, 2006; R. Cuadros, The sH stress trajectories were established by connecting,
point by point, the orientation of the inferred direction in all
N grid cells. An example of a portion of the stage H3 trajectories
from the central part of the deposit is shown in Figure 9 and
the trajectories for all alteration stages in Figure 10.
The local horizontal principal stress trajectories (sH) estab-
W E lished for all alteration events (propylitic to late hydrothermal;
Fig. 10) are nonlinear and noncoaxial and display dissimilar
orientations through the successive events and within the dif-
ferent very important fault-bounded panels of the mine. They
S differ from those of the previous and following events, and
Without in almost all cases the local sH trajectory does not coincide
correction
with the approximately east-northeast constant orientation
ch

of the tectonic far-field stress direction for the time of the


N
Ben

3
W E sh.fr sh.fr

e-s e-s
ext ext
S
e-s e-s
With correction
1
Fig. 7. Application of the Terzaghi correction to a rose diagram. The diagram sh.fr sh.fr
corresponds to an example of the structures before and after the correction.
This correction was done in all diagrams and consists of a geometric tech-
nique that modifies linear and planar samplings of fractures with respect to Fig. 8. Expected orientation with respect to the principal stress axes of new-
directional bias, determining a more precise orientation of the structure with formed shear (sh.fr), extensional shear (e-s), and extension fractures (ext).
respect to the rock slope direction (Terzaghi, 1965). From Sibson (2000).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
14
Table 3. Structural Classification and Very Important Fault Characterization (see Figs. 3, 10)

by Carleton University user


Fault zone Structure Core Damage zone Breccia and vein textures Crosscutting relationships Permeability/barrier potential

Americana Heterogeneous; filled Multiple (>1 m) of dark-gray Nearby extension vein sets Breccias vary from jigsaw Veins crosscut the fault’s Bounds to the east the
  fractures in damage and   gouge containing quartz   formed by quartz-Mo dm-   puzzle textures and can vary   fabric and are locally cut by   advanced argillic alteration
  localized deformation zones   crystals, banded enargite,   thick breccias, cracks, and   from angular to subrounded   the fault; brecciated, early   (LH; Figs. 3, 4, 10); conduit
  and sphalerite veins   gouge, all semiparallel to   quartz-Mo fragments, in   quartz fragments float on a   for the quartz-Mo
  fault traces that transect and   some cases cut by later   hydrothermal matrix with
  crosscut quartz-Mo breccias   quartz-pyrite veins   fragments of quartz-
  molybdenite veins within
  late quartz-sericite breccias
  (Faunes et al., 2005)

NNW Heterogeneous structural Multiple (0.5–1 m thick), Formed by semiparallel Quartz-Mo veins distribute Overprints early potassic Alteration and mineralization
  fabric   with dark-gray gouge;   fault traces, veins, breccias,   parallel and adjacent to   related Cu-sulfide   are not limited by this fault
  fault veins and fault zones   and dm-thick gauge zones;   major structures   mineralization.   (Figs. 6, 10); conduit for
  mineralized with enargite-   closely associated with   quartz-Mo and enargite
  sphalerite ± pyrite and   extension veins   stages
  minor coarse-grained
 covellite-digenite

EBN Splits into multiple Core zone formed by single- Closely associated with Dilation breccias that Displaces the Chuquicamata Controls the distribution of
  mineralized strands, some   state breccia alignments   reciprocal crosscutting   display quartz-pyrite-Cu   Intrusive Complex by less   alteration and mineralization
  of them bending to become   relationships between fault   sulfide veins and sericite   than 100 m (see also   in early and late stages, as
  parallel to the Americana   segments and extension   alteration selvages, as   Lindsay et al., 1995;   reflected by the orientation
  fault zone (Perry, 1952;   vein sets, with common   well as later clay gouge   Rivera et al., 2012; Fig. 2);   and distribution of quartz-
  Lindsay et al., 1995)   alteration-mineral infill   reciprocal crosscutting   sericite and advanced argillic
  relationships with extension   alteration stages
  veins
JORGE SKARMETA

EBS Subsidiary parallel Branches of the system in the Closely associated with Breaching and cementation Extends outside the pit Bounds the distribution of
  structures exert control on   pit present quartz-pyrite-Cu   extension veins   histories with rebrecciated   where it makes minor   alteration and mineralization
  superficial geometry of the   and sericite alteration along   breccias and late stylolite   displacements (<50 m) on   in late quartz-sericite and
  Banco porphyry (Figs. 2, 3)   the margins, as well as later   textures   Mesozoic and Paleozoic   advanced argillic alteration
  clay gouge and chlorite-   units (Fig. 2); reciprocal   stages
  filled fractures     crosscutting relationships

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


  with extension veins

Zaragoza Veins of this system were Wide damage zone with Impervious behavior at the
  historically mined   massive pyrite veins in the   principal sericite event
  upper benches that were   and most of the main
  superimposed over quartz-   hydrothermal stage; no
  Mo veins at lower benches   C-type veins have been
  recognized to the west of it
  (Figs. 3, 4, 10)

C2 Silicified along strike, Associated with Controls on the emplacement Impervious behavior at the
  associated with quartz   extension veins   of the Chuquicamata   principal sericite event
  veins   Intrusive Complex and the
  hypogene and supergene
  mineralization

EBN and EBS are Estanques Blancos North and South


STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 15

(tectonic), erosional, magmatic, and thermal (Fig. 10d-g). The


thermal stresses are dependent on the magma and/or hydro-
thermal fluid’s temperature (T) at the corresponding depth (z)
(Figs. 4, 10). Temperature and depth variables were assigned
from paragenetic temperature-pressure equilibrium phase
N relationships (e.g., Seedorff et al., 2005; Richards, 2011), fluid
inclusion trapping temperatures determined at Chuquica-
mata (S. Collao, unpub. report, 1987; Vega, 1989; Lindsay,
1998; Arnott, 2003), and comparisons with equivalent well-
studied porphyry copper systems (cf. Titley, 1993; Rusk et al.,
2004; Masterman et al., 2005; Seedorff et al., 2005; Sillitoe,
2010; Monecke et al., 2018).

Resultant Stresses and Overpressure Zonation


During Alteration
Stress partitioning in time
At any depth and associated alteration stage (propylitic-late
hydrothermal; Fig. 4; Table 1) of porphyry emplacement,
the total horizontal stress, sTH, corresponds to the sum of the
partial stress components: elastic-uniaxial (sH) thermal (sT),
magmatic, erosional, and tectonic (st), equivalent to the sum
of equations (3) and (4), starting from the commonly accepted
~6 km from surface (Seedorff et al., 2008; Proffett, 2009; Sil-
litoe, 2010; Fig. 3):
sTH ≤ sH + sT + st
         Dz
[ ]
= —– ⋅ (1 – 2v)ρ ⋅ g + a ⋅ E ⋅ DT + 3 ⋅ sV. (5)
        1–v

The synthesis of the stress components (sH, sT, st, sHT; eq.
5) with depth (z) illustrates the role played by thermal and
elastic stresses during the evolution of the deposit (Fig. 12).
100m Repeated magmatism and fluid activity generate significant
heat energy that contributes to large stress variations (Fig.
Fig. 9. Example of the determination of the principal horizontal stress tra- 13). Uplift and consequent erosion makes the overburden
jectories from the central part of the deposit in a gridded portion of the H3 load decrease, contributing to a reduction of the elastic-
veins. Based on Terzaghi corrected rose diagrams (Fig. 7). related horizontal stress (eq. 1). Note that the overall trend
of the stress calculations remains unchanged when adapted
Incaic deformation phase (~40–30 Ma; Pardo-Casas and Mol- for different values of critical temperature, depth, and rock
nar, 1987; Somoza, 1998; Somoza and Ghidella, 2005; Sdro- parameters.
lias and Muller, 2006; Chen et al., 2019; Figs. 10, 11). The The resultant total horizontal stress varied during emplace-
variable q angle that the principal horizontal stress trajectory ment, from positive (~400 MPa) at the magmatic and potassic
forms with segments of any particular fault will condition the stages (~6 km), gradually decreasing to reduced, close-to-zero
local shear and normal stress magnitudes (see Figs. 12, 13). stresses (Figs. 12, 13) at the main hydrothermal, neutral con-
Substantial differences in stress orientation can be visualized traction-extension transition (i.e., 2–3.5 km), to finally becom-
when the stress trajectories from all events are plotted one ing extensional at shallow late-hydrothermal levels (<2 km),
on top of each other, as if they had occurred simultaneously where sV > sH.
and not on superimposed stages, even if developed with time
hiatuses between them (Fig. 10h)—the exception being the Differential stress
stage H1 trajectories that are relatively straight and contin- The differential stress (s1 – s3) at every depth and tempera-
uous, although not parallel to the extant (~33 Ma) far-field ture was estimated from the variations of the total horizontal
stress direction (Figs. 10c, 11). stress (sTH) and of the vertical stress (sV) through time. When
Distortions from a linear sH trajectory have been related to s1 = sTH and s3 = sV the tectonic regime will be compressive
local parameters, such as anisotropies (cf. Ojala et al., 1993; (sH > sV), whereas when s1 = sV and s3 = sTH (sV > sH) will
Healy, 2009), overpressured faults (Jaeger and Cook, 1979), be extensional (Figs. 12, 13). Note, however, that at the stress
damage-induced weakness in faults (Faulkner et al., 2006), changeover, a wrench state of stress (s2 = sV) will develop
and variable depth- and temperature-dependent local compo- transiently.
nents of the total horizontal stress. The total horizontal stress The calculated differential stresses involved in the forma-
consists of the sum of all stress constituents, including far-field tion of Chuquicamata progressed from a maxima, at the initial

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
16

by Carleton University user


JORGE SKARMETA

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


Fig. 10. (a-h) Stress trajectories per alteration stage (propylitic-late hydrothermal) plotted over the very important fault skeleton. The sH stress orientation determined at
every particular grid was point-by-point traced along the area with sufficient data. See example in Figure 9. Figure 10h shows the differences in principal stress orientation
when the stress trajectories from all events are plotted one on top of the other, as if they had occurred simultaneously and not successively. Alteration stages abbreviated
and color coded as in Figure 4.
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 17

late-magmatic, background potassic, and intense potassic


stages (background potassic to early hydrothermal), in which
sH > sV to a minima at the final late-hydrothermal stages, flip-
10 S
South ping from compression (sH > sV) to extension (sV > sH), at the
America main hydrothermal stage (H2-H4; Fig. 4) where differential
Plate stresses become neutral or near isotropic at the contraction-
Present extension stress inversion. This stage developes when sH ≈ sV
30 20 10
and coincides with stockwork formation, significant tempera-
40 Ma Nazca Plate
20 S
ture variations and with the largest volumes of copper added
50 to the system (Figs. 4, 12, 13). A late extensional stress devel-
60 Incaic Event opment (late hydrothermal) occurred in a relatively short
70 35 Ma 30 Ma period of time, between 32 and 31 Ma (Ballard et al., 2001;
Barra et al., 2013; Fig. 13). This late stress configuration favors
20 10
30 the development of new vertical extension hydrofractures
40 40 S
(Ridley, 1993) and normal fault reactivation, all associated
50 with decompression, gravitational collapses (e.g., Vanderhae-
59 Ma
60 ghe and Teyssier, 2001; Masterman et al., 2005; Maydagán et
90 W 70 W
al., 2015), and rising of hydrothermal fluids associated with
massive hydraulic discharges (Fournier, 1999), together with
Fig. 11. Reconstruction of plate convergence vectors for the Nazca plate, rel-
ative to South America, during the Cenozoic, representing the far-field stress
the formation of late veins and breccia bodies, such as those
orientation (redrawn from Pardo-Casas and Molnar, 1987, and Maksaev and associated with the Americana fault (Ossandon et al., 2001;
Zentilli, 1999). Faunes et al., 2005; Figs. 3, 14; Table 3).

LH H4 H3 H2 H1 EH PR BK
180-220 220-270 300 550-500 ~ 550 550-400 ~ 450 550-650 T° (C°)

Background
Propylitic
Intense
Qz - Mo
Principal
Argillic
QS
Advanced

500

400
Frictional faulting
300 limit
Total horizontal stress
Argillic

200
Stress (Mpa)

Vertical stress

100

0
Potassic
Sericite

Potassic

Elastic, posterosion
horizontal stress
-100 Differential stress

-200
Alteration

-300
Thermal stress
-400 Depth
(Km)
0 1 2 3 4 5 6
31 33 33,5 Age (Ma)
Fig. 12. Representation of the partitioned stress components of Chuquicamata, plotted against time and depth (color codes
as in Fig. 4). Curves: dashed gray represents the maximum tectonic stress indicated by the upper crust’s limiting stress of
preexisting faults (st, eq. 3; from Jaeger and Cook, 1979); solid blue line shows the total horizontal stress (sTH, eq. 5); solid
red line represents the elastic or uniaxial stress (eq. 1), solid purple line represents the vertical stress (sV); solid green line is
the thermal stress (sT, eq. 4), and the solid yellow line represents the differential stress (s1 – s3). Abbreviations: QS = quartz-
sericite, Qz =quartz. Alteration stages abbreviated and color coded as in Figure 4.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
18 JORGE SKARMETA

LH H4 H3 H2 H1 EH PR BK Magmatic and porphyry domains

New vertical fractures Low


differential
and
deviatoric
stress

Initial state
of stress

T0
Zone of stress reversal

Fig. 13. Mohr diagram constructed with stress data determined for Chuquicamata (Fig. 12). Differential stresses calculated
from maximum horizontal and vertical stresses at every alteration stage and depth (see Figs. 4, 12). The composite range of
failure for intact rock (gray area) indicated by solid black lines, and reshear condition for a cohesionless fault (dashed black
line) plotted on a shear stress, t, against effective normal stress, sn, field. Critical stress circles are shown for the seven altera-
tion stages indicated with the same color code as Figure 4. The red dashed lines separate the contractional from extensional
domains and cross at the point of stress reversal. Note that newly formed fractures will only develop when the stresses are
sufficient to intersect the intact rock envelope, principally at the extensional domain.

Stress orientation and vein mineralization as faults truncate veins and veins crosscut faults with a simi-
Vein sets are considered to develop at high angles to the lar hydrothermal infill (Figs. 6, 10; Table 3). The close asso-
local least principal stress axis (Sibson, 2001); hence, the ciation between the Americana, NW, and Estanques Blancos
vein rotation observed (Fig. 10) requires the least principal North and South fault segments with significant popula-
stress orientation to change several times during the life of tions of nearby extension vein sets are indicative of hybrid-
the relatively short duration hydrothermal processes (~2 m.y.; type fault-vein dilatational meshes (cf. Hill, 1977). In the
Table 1). The veins in Chuquicamata do not show orienta- H1 alteration stage the B-type veins cut across most of the
tions consistent with one single stress field, and they change faults, whereas in subsequent H3 and H4 alteration stages the
within and between the different paragenetic stages (Figs. 4, veins are limited by the faults (Fig. 10c). Low-displacement
10). This means that the local stress trajectories do not mimic fractures developed in conjunction with high-dilation exten-
the orientation of the regional stress field—an assumption sion veins (cf. Newhouse, 1940; McKinstry, 1948) correspond
commonly used to establish the main structural controls in to hybrid structures (Sibson, 2000, 2001; Cox, 2010, 2020),
Chuquicamata (e.g., Reutter et al., 1991, 1996). The way in which develop in low or isotropic differential stress condi-
which the orientation of the veins, controlled by the stress tra- tions, where the two principal horizontal stresses may be
jectories, vary during the deposit’s timescale indicates that the interchanged (i.e., Tosdal and Dilles, 2020).
far-field tectonic stress is not the principal component of the The reciprocal, crosscutting relationships between faults
local stress budget. and veins provide insights into the fault’s hydraulic behavior at
the different stages of mineralization. The changes in the frac-
Fault-vein relationships ture type, from hybrid faults to sheeted veins to stockworks,
Mutual crosscutting relationships between different veins reflect variations in differential stresses and fluid pressure
and very important faults suggest synchronous development, conditions and reveal the differences in the magnitude of the

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 19

H> V v V> H

V
Erosion and surface
Degradation V> Pf > H
Pf < H

Brittle

Ductile
H
H
H H
~ 6 Km
~ 3-4 Km
Ductile-brittle
transition
~ 400°C

H> Pf > V

Alteration
Quartz-Alunite
Intrusions Quartz-Prophylite
Postmineral Sericite
Synmineral Sericite - Clorite
Fluid - flow Magma
Premineral Potassic / Propylitic

Fig. 14. Schematic time slices of the telescoping process of Chuquicamata showing the evolution of the main magmatic,
fluid, and alteration-mineralization types during progressive magma cooling and paleosurface degradation. Left: syntectonic
high-temperature magma emplacement (>400°C; Fournier, 1991, 1999) that under contractional stresses generates sealing
carapaces that isolate the lower, lithostatically pressured ductile panel from an upper brittle hydrostatically pressured one.
At the contraction to extension, stress reversal (Figs. 12, 13) stresses become neutral, quasi isotropic, where fluid pressures
substantially increase, and multiple orientation stockworks develop. Right: As magma solidifies and the entire system cools
below ~400°C, the rock can fracture in a brittle fashion (Fournier, 1999); once the extension stress regime is established,
seal breaching can be attained by downward propagation of normal faults and extension fractures. Lithostatic pressure gives
way to hydrostatic pressure and erosion progressively degrades the paleosurface. Massive sulfide breccias crosscut previous
alteration stages. Based on Sillitoe (2010) and Fournier (1999).

local principal stress. Recementation of well-oriented fault Burov, 2003) and is destroyed by resealing by mineral infill
segments (Sheldon and Micklethwaite, 2007) can allow them and vein formation (Sibson, 2001; Cox, 2020; Tosdal and
to reach an impermeable condition and therefore modify their Dilles, 2020; Figs. 4, 10). The generation of veins and their
mechanical response to the applied local stress, in comparison crosscutting relationships, in which the older vein sets are
with nonrecemented segments. These contrasting hydraulic crosscut by subsequent younger ones (Fig. 5), show that, in
behaviors can be related to at least two independent factors: Chuquicamata, permeability was transiently created and epi-
(1) the local orientation of the principal stress with respect to sodically destroyed by resealing through mineral precipitation.
the fault orientation (the angle q), visualized by the sH stress Therefore, permeability development was a time-controlled
trajectories (Fig 10), and (2) the variation in the magnitude dynamic competition between permeability enhancement
of the principal stress, visualized by means of the variation and destruction, in which vein-forming processes led to the
in the differential stress (Fig. 12). The very important fault’s recovery of the original cohesive strength of the rock mass.
reactivation capacity is therefore variable in space and time, Breached segments of the Americana fault with brecciated,
according to the local stress field and the local angle q, depen- early quartz fragments floating in a hydrothermal cement with
dent upon the stress trajectories (Figs. 10, 12, 13). fragments of quartz-molybdenite veins within late quartz-ser-
icite breccias (Table 3; Faunes et al., 2005) provide evidence
Permeability cyclicity of breaching and cementation processes in which faults regain
Hydrofracture permeability is created by episodic magmatic, cohesive strength enabling overpressure recuperation (cf.
hydrothermal, and/or tectonic events (Guillou-Frottier and Nguyen et al., 1998; Sibson, 2000, 2004; Cox, 2010). However,

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
20 JORGE SKARMETA

in fault-related breccia systems developed around misori- under high strain rates and switching stress regimes (from
ented fault strands, such as parts of the Estanques Blancos sV = s3 to sV = s1; Figs. 12, 13), will enable the downward
North fault (Fig. 3), resealing leading to an impermeable con- propagation of ruptures and shears to breach seals, inducing
dition may make the rock become harder (reseal hardening; multiple episodes of fault-valve discharge (Sibson, 2004; Cox,
Woodcock et al., 2007), favoring a subsequent mineral phase 1995) and drainage of overpressured hydrothermal fluids and
to develop different core and damage zones in relatively formation of late-hydrothermal breccia and mineralized frac-
weaker intact rock (Table 3). tures (Fig. 14). The mechanism to trigger fault propagation
may be provided by an increase of the differential stress (after
Vertical overpressure zonation cooling and erosion; see events H3 to late hydrothermal in
The presence of high-dilation extensional and extensional- Figs. 12, 13) to an amount sufficient to allow normal reacti-
shear mineralized veins (Fig. 4; Table 1) and stockwork vation of favorably oriented fault segments, accompanied by
networks along and adjacent to existing misoriented very fault-fracture drainage, pressure drop, abundant quartz pre-
important faults (Table 3) without evidence of shear displace- cipitation (Fournier, 1999; Rusk and Reed, 2002), and fluid
ment suggests the development of significant overpressure redistribution (Sibson, 2004). Note that the distribution of
and low differential stress (Etheridge, 1983; Cox, 1995, 2010, alteration stages at Chuquicamata shows a slight northward
2020). Overpressure can be sustained provided no throughgo- tilt as the lower-temperature H3 associations dominate in that
ing favorably oriented faults occur, which, if present, would direction (Fig. 6), suggesting a down-to-the north displace-
be preferentially reactivated (Sibson, 1990, 2012; Cox et al., ment along the northeast faults (e.g., Estanques Blancos
2001; Cox, 2005, 2010). The reactivation of severely misori- North and South; Fig. 6h) after or during the H4 and late-
ented faults under varying stress fields (Figs. 12, 13) provides hydrothermal events.
key evidence for predominantly fluid driven failure at low dif-
ferential stress in overpressured systems (Pf > sH; Cox, 2010). Discussion
The generation and maintenance of lithostatically overpres- Chuquicamata is a magmatic-hydrothermal system developed
surized vein-fracture meshes (i.e., the main hydrothermal at variable interrelationships between the local stresses and
events, H1 to H4 in Fig. 4), require a transient isolation from hydrothermal fluids. Important controlling factors are stress
the overlying, hydrostatically pressured, late-hydrothermal magnitudes and orientations, fluid overpressure, fracture
extensional regime. This isolation can be attained by rheo- development and sealing, vein formation at succeeding pulses
logical seals caused by the development of igneous carapaces of hydrothermal fluids of variable temperature, internal
at the ductile-brittle transition (Burnham, 1985; Fournier, fault paneling at different overpressures, and an active mag-
1999) or by relict bands of relatively more impermeable folia- matic arc that was progressively uplifted and eroded during
tions that may represent synemplacement fabrics related to mineralization.
the intrusion of various phases of the Fortuna Intrusive Com-
plex and Chuquicamata Intrusive Complex (Lindsay, 1998; Stress regimes and porphyry emplacement
Faunes et al., 2005; J. Dilles, unpub. report, 2008; Rivera et Arc polyphase magmas and porphyry copper deposits
al., 2012). are linked to unusual regional-scale geodynamic settings
A schematic vertical section through the Chuquicamata (Mpodozis and Cornejo, 2012) favorable for mineralization
magmatic-hydrothermal system can be separated in three (Sillitoe, 2010). They are short transient periods in geologic
broad structural levels with characteristic alteration, fluid time (Blundel, 2002; Sillitoe and Perrello, 2005) that imply
pressure, and stress distributions (Seedorff et al., 2008; Prof- some fundamental physical processes (McCuaig and Hronsky,
fett, 2009; Sillitoe, 2010). These correspond to (1) a high- 2014).
temperature (>500°C), deep (~6–~4 km), potassic alteration There is a relative agreement that porphyries develop along
developed in a contractional stress regime, (2) an overlying subduction or collisional contractional plate margins (Silli-
overpressurized hydrothermal system, developed at interme- toe, 1973, 1997; Tosdal and Richards, 2001; Richards, 2003;
diate temperatures (~375°–400°C) and depths (~1.5–4 km), Mpodozis and Cornejo, 2012); however, it has also been pro-
that is isolated from the uppermost rock column by temper- posed they can occur in weak contractional to near-neutral
ature-generated impervious seals (e.g., Fournier, 1999), and tectonic environments developed in response to bulk stress
(3) a late-hydrothermal, shallow depth (<1.5 km), hydrostati- changes (Corbett and Leach, 1998; Maksaev and Zentilli,
cally pressurized domain emplaced in an extensional stress 1999; Richards et al., 2001; Tosdal and Richards, 2001; Cooke
regime, where an unconfined column of fluids circulated et al., 2005; Richards, 2009, 2013; Kloppenberg et al., 2010;
through the uppermost brittle cap rocks. At the end of the Bertrand et al., 2014; McCuaig and Hronsky, 2014). At the
process, progressive uplift and erosion cause all three levels deposit scale, neutral stresses may be locally and transiently
to become juxtaposed, at a vertically condensed, telescoped achieved by internal stress accommodations that result from
structural level (Fig. 14), in which the deep potassic and the the variations of the partial stress components during the devel-
KSil (see Table 1) are overprinted by the late-hydrothermal opment of the deposit, independent from the far-field stress
high-sulfidation assemblages (Faunes et al., 2005). even when it remains unchanged. Porphyry copper deposits
The intermediate vein-fracture meshes and stockworks are generally emplaced in contractional plate margins where
indicate conditions of an overpressurized hydrothermal sys- high-angle fault reactivation, tectonic inversion, and adakite-
tem. Magma upwelling may pool fluids into shallower depths like magmatism develop (Thieblemont et al., 1997; Gutscher
below the impermeable rock cupola (cf. Cloos and Sapiie, et al., 2000; Kay and Mpodozis, 2001; Oncken et al., 2006;
2013), generating anomalous fluid-pressure gradients that, Loucks, 2014). The inversion of the severely misoriented

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 21

high-angle faults (>60°–70°; Sibson, 1985), commonly associ- while Tosdal and Richards (2001) and Gruen et al. (2010) sug-
ated with porphyry systems (McClay et al., 2002; Amilibia et gest that veins and fractures result from the overlapping of
al., 2008), is fundamentally related to the high stress ratio and renewed magma intrusions and externally imposed tectonic
fluid overpressures (Sibson, 1995; Haseguawa, 2017) required stresses. Tosdal and Dilles (2020) indicate that in porphyry
to overcome the higher frictional constraints that result from environments two vein orientations are commonly seen. The
an intensification of the deep-seated slab-flattening tectonics principal one is generally formed by closely spaced sheeted
developing at the time (Mpodozis and Cornejo, 2012, and ref- veins that correlate with the extant far-field stress orienta-
erences therein). tion at the intrusion time. Subsidiary vein sets, which may be
orthogonal to the main system, develop radially or concen-
Stress fields trically and reflect the localized effect of magma intrusion,
Heterogeneities that make the local stress field depart from hydrofracturing, and host-rock dilation, setting an anisotropic
the far-field orientation at the camp scale have been pointed fabric that is subsequently used by later vein systems, hence
out as controlling factors for deposit formation (e.g., Dubé et defining the principal vein direction at an early stage in por-
al., 1989; Ojala et al., 1993; Faulkner et al., 2006). In Chuqui- phyry development. However, most porphyries commonly
camata, the prevailing local principal horizontal stress orien- exhibit multiple vein sets that develop at different sequenced
tation, at the majority of the alteration stages, departs from paragenetic stages (cf. Gustafson and Hunt, 1975; Titley et al.,
coaxiality and does not match the orientation of the far-field 1986; Titley, 1993; Stanley et al., 1995) in which the constant
stress (compare Figs. 10, 11) commonly accepted from plate angular relationships have been rotated. At Chuquicamata,
convergence vectors (cf. Pardo-Casas and Molnar, 1987; the vein orientations (Fig. 10) show high variability between
Somoza, 1998; Somoza and Ghidella, 2005). Given the over- the different alteration stages, implying that the local prin-
whelming nonlinearity of the stress trajectories, the angle q cipal stress orientation varied significantly throughout the
sustained between the maximum local horizontal stress and life of the relatively short duration hydrothermal processes
fault segments will only be locally well oriented for optimal (~2 m.y.; see Hypogene Alteration section, Table 1). The lack
reactivation (Sibson, 2017). of a hereditary behavior can be better visualized in Figure
The close association and reciprocal crosscutting relation- 10h, where all stage stress trajectories are plotted one on top
ships between fault segments with nearby extension vein sets of each other and in which a preferential direction cannot be
with common alteration-mineral infill (e.g., Americana, NW, recognized. Implicit in the argument is that vein formation
and Estanques Blancos North and South faults; Fig. 4; Table was followed by an effective fracture healing process, devel-
3) are indicative of hybrid-type fault-vein dilatational meshes oped during and in between different alteration stages, when
(cf. Hill, 1977; Cox, 1995). rocks had effectively recovered their homogeneity.
Very important faults and spatial distribution of alteration Fluid pressure evolution
The spatial distribution of the alteration associations (propy- At the compressional state, overpressure is commonly cap-
litic-late hydrothermal; Figs. 4, 6) within the very important tured and maintained in low differential stress small-scale
fault framework (Fig. 6h) indicates that major faults had dif- veins with a homogeneous fluid distribution throughout the
ferent hydraulic behaviors throughout the hydrothermal his- fracture network (e.g., Sanderson and Zhang, 1999; Figs. 12,
tory (cf. Byerlee, 1993). Whether a very important fault acted 13). It has, however, been pointed out (e.g., Byerlee, 1993)
as a localized distribution conduit, barrier, or a combination that preexisting very important fault-type sealed faults may
of both will depend on the relative percentage of fault core compartmentalize high-pressure fluids, making the fluid dis-
and damage zone structures (Caine et al., 1996; Faulkner et tribution relatively heterogeneous. The stress switching, from
al., 2006, 2010; Sheldon and Micklethwaite, 2007). Their rela- horizontal compression to extension, favors fluid release (and
tive proportions represent a finite stage of a time-dependent reduction of fluid pressure), commonly accompanied by the
incremental process, as the hydraulic behavior of faults was development of vertical extension fractures (Sibson, 2004;
subject to modifications over multiple strain and fluid flow Van Noten et al., 2011, 2012), fault ruptures and seal breach-
alteration-associated events. Such events may reduce perme- ing (Fournier, 1991), overpressure homogenization, and
ability within fault cores as they regain cohesive strength by stress variations (Cox, 2010, 2020; Sibson, 2017). Late veins
hydrothermal recementation (e.g., Barton et al., 1995; Sibson, and breccia bodies contained in the Americana fault system
2000; Cox, 2005, 2010, 2020; Micklethwaite et al., 2010). The (Ossandon et al., 2001; Faunes et al., 2005; Figs. 3, 14; Table
relogged very important fault intercepts (Tables 2, 3) reflect 3) are associated with vertical extension fractures and massive
the degree to which the processes of strain localization versus hydraulic discharge of hydrothermal fluids (Fournier, 1999;
strain distribution compete as the fault zone and the altera- Vanderhaeghe and Teyssier, 2001).
tion are modified by ambient conditions (Couples and Lewis, The evolution of the fluid pressure (Pf) can be visualized
2007). by means of a brittle failure mode plot (BFM; Sibson, 1998,
2000) that defines the decrease of the differential stress with
Vein orientation increasing overpressure at different effective vertical stress,
Veins and mineralized fractures in porphyry Cu deposits show s'V (where s'V = sV – Pf), during stress inversion. Maximum
substantial variations in orientation and structural style with sustainable overpressure is likely to develop at the low differ-
respect to their inferred depth of formation (e.g., Wilkinson ential stress compression-extension transition (Sibson, 2004;
et al., 1982). Heidrick and Titley (1982) proposed that veins Van Noten et al., 2012). Conditions for the development and
have a vertical zonation related to shallow subvolcanic stocks, maintenance of overpressures (Pf > sH) are those required to

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
22 JORGE SKARMETA

generate new hydraulic fractures in intact rock (i.e., stock- internal geologic processes. The tectonic evolution of the
works) or for the dilation of existing multiple-orientated very continental margin during the Incaic phase of deformation
important faults, likely to form at small stress differences is ultimately associated with a high degree of coupling and
(Jolly and Sanderson, 1995, 1997; Stephens et al., 2004) and consequent orogenic shortening, adakite-like magmatism
high stress R ratios (Delaney et al., 1986; Baer et al., 1994), linked with a flat-slab subduction regime (e.g., Mpodozis and
a relationship that incorporates the principal stresses, pore- Cornejo, 2012). The shortening of the margin results in a
fluid factor (lv), and orientation of preexisting very important thickened arc, highly pressurized lower crust magma cham-
fault anisotropies (Healy, 2012). bers, and magma with increased residence time that contrib-
utes to extreme geochemical fractionation. The far-field stress
Permeability evolution at the porphyry emplacement time had a constant approxi-
The repeated regeneration of fracture permeability becomes mately east-northeast linear orientation (Fig. 11), which
a requisite to sustain the extreme high fluid fluxes associated contrasts with the observed internal stress directions at the
with porphyry formation. The cyclical incremental infill in deposit scale (Fig. 10). Therefore, nontectonic stresses need
extension and extensional shear is commonly recorded in vein to be considered to explain the stress trajectory variations
textures that suggest that host fractures were locally gapping throughout the alteration history. Consequently, stresses must
at the time of hydrothermal deposition (Ramsay, 1980; Boul- be linked with depth and time-related variables, such as over-
lier and Robert, 1992; Cox, 1995, 2005; Bons et al., 2012). burden weight and temperature-related stress.
Quartz-Mo blue veins and late enargite veins from Chuquica- A composed section through Chuquicamata shows that
mata (Lindsay, 1998) reveal multigenerational cyclical growth the stresses progressed from a deep initial, lithostatically
histories that result from recurrent fluid pressure oscillations pressurized contraction into a late, shallower hydrostatically
and sequential variation of silica solubility in hydrothermal pressurized extension regime (Figs. 12–14). The contraction-
fluids in the pressure and temperature ranges of porphyry- extension transition postdates the early high-temperature
related mineralization (i.e., Rusk and Reed, 2002; Rusk et A and B quartz veins. It coincides with the most significant
al., 2008; Monecke et al., 2018). Recurrent breaching and mineralization influx at an overpressurized, quasi isotropic
cementation processes on fault segments, as brecciated, early differential stress regime that is isolated from the uppermost
quartz fragments floating in a hydrothermal matrix, and frag- rock column by impervious seals. The stress-fluid conditions
ments of quartz-molybdenite veins within late quartz-sericite throughout the development of the porphyry provide bound-
breccias have been described in the Americana fault (Table 3; ary conditions for fracturing, veining, and mineralized or bar-
Faunes et al., 2005). ren fault development.
As the rates of fluid pressurization generally exceed those Heat energy released by repeated intrusions and elastic
of tectonic loading (Micklethwaite and Cox, 2004, 2006; Ste- stresses developed during uplift and erosion drive continu-
phens et al., 2004; Tenthorey and Cox, 2006) and magma ous bulk stress readjustments and stress-fluid interactions.
supply rates (Tibaldi, 2015; van Wyk de Vriers and van Wyk The combined factors may trigger local fault reactivation,
de Vriers, 2018), the recovery of cohesive strength devel- breaching of self-sealed zones, damage zone modification,
ops on the timescale of rupture-recurrence (Cox, 2016) and brecciation, permeability increase/decrease, and discharge
dike-driven intrusion rates (Petford, 1996). Heat conduc- and focusing of hydrothermal fluids. The overall local stress
tion models developed at the Butte porphyry (Mercer et al., deviations from the far-field orientation (Figs. 10, 11) indicate
2015) show that vein formation and dike injection formed by that the approach of constant stress magnitude or orientation
short-lived episodes of hydrofracturing in the range of tens of versus dilatational space, frequently utilized for predicting ore
years. The timescales for the formation and cooling of vari- deposit geometry and emplacement, does not apply to camp-
ous generations of hydrothermal quartz veins range from tens scale porphyry systems governed by time-fluctuating stresses
to tens of thousands of years, which in geologic times can be (Fig. 10).
regarded as effectively instantaneous. This is supported by
microstructural and experimental evidence in which high- Acknowledgments
permeability microcracks heal at geologically short timescales This work was developed while I worked in the Corporate
(Brantley et al., 1990; Morrow et al., 2001; Zhang et al., 2001; Exploration Group of Codelco. The founding of the work
Tenthorey et al., 2003). Modern examples of vein arrays and was provided by the Corporate Exploration Group and by
associated fault rocks from fluid-active high-temperature self- the Geology Super-Intendancy of Chuquicamata. Both are
sealing regimes develop in timescales equivalent to those of acknowledged for the support and permission to publish the
the driving structural process (Blanpied et al., 1992; Lowell et results of the study.
al., 1993; Hickman et al., 1995; Olsen et al., 1998; Cox 1999, Very insightful and helpful criticism to this manuscript,
2005, 2010, 2020). Analogue modeling and measurements in which significantly improved its content, clarity, and pre-
volcanic-geothermal upflow vents along fracture zones dem- sentation, was provided by Greg Corbett, Roberto Freraut,
onstrate that sealing by quartz precipitation at high tempera- Jon Hronsky, Carlos Huete, Darryl Lindsay, Victor Maksaev,
tures occurs on timescales of decades (cf. Moore et al., 1994). Ken McClay, Rodrigo Morel, Constantino Mpodozis, Carlos
Munchmeyer, Bob Pankhurst, Jose Piquer, and Juan Carlos
Summary and Conclusions Toro. Robert Monypenny is thanked for an exhaustive revi-
The Cu mineralization at Chuquicamata, as in many other sion of the English. Dick Tosdal and an anonymous reviewer
giant porphyry deposits, is the result of time and spatial coin- are thanked for valuable suggestions that helped to organize
cidences and interactions of several external tectonic and and clarify the presentation. Thorough manuscript reviews

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 23

and helpful suggestions by Associate Editor Shaun Barker Bons, P., Elburg, M.A., and Gomez-Rivas, E., 2012, A review of the formation
are much appreciated. Special thanks go to Hector Barbage- of tectonic veins and their microstructures: Journal of Structural Geology,
v. 43, p. 33–62.
lata, who diligently, and with utmost patience, drafted and Boric, R., Díaz, F., and Maksaev, V., 1990, Geología y yacimientos metalífe-
redrafted most of the figures included in the text. ros de la región de Antofagasta: Servicio Nacional de Geología y Minería,
Boletín 40, 246 p.
REFERENCES Boullier, A.M., and Robert, F., 1992, Paleoseismic events recorded in
Álvarez, O., and Flores, R., 1985, Alteración y mineralización hipógena en Archaean gold-quartz vein networks, Val d’Or, Quebec, Canada: Journal of
el yacimiento de Chuquicamata, Chile: Actas del IV Congreso Geológico Structural Geology, v. 14, p. 161–169.
Chileno, Antofagasta, 1985, Proceedings, v. 2, p. 78–100. Brantley, S.L., Evans, B., Hickman, S.H., and Crerar, D.A., 1990, Heal-
Ambrus, J., 1979, Emplazamiento y mineralización de los pórfidos cuprífe- ing of microcracks in quartz—implications for fluid flow: Geology, v. 18,
ros de Chile: Unpublished Ph.D. thesis, Salamanca, Spain, Universidad de p. 136–139.
Salamanca, 308 p. Braxton, D.P., Cooke, D.R., Ignacio, A.M., and Waters, P.J., 2018, Geology
Amilibia, A., and Skarmeta, J., 2003, La inversión tectónica de la Cordillera of the Boyongan and Bayugo porphyry Cu-Au deposits: An emerging por-
de Domeyko en el Norte de Chile y su relación con la intrusión de sistemas phyry district in northeast Mindano, Philippines: Economic Geology, v. 113,
porfídicos de Cu-Mo: Actas del 10° Congreso Geológico Chileno, Concep- p. 81–131.
ción, October 6–10, 2003, Proceedings, CD. Burnham, C.W., 1985, Energy release in subvolcanic environments: Implica-
Amilibia, A., Sabat, F., McClay, K.R., Muñoz, J.A., and Chong, G., 2008, tions for breccia formation: Economic Geology, v. 80, p. 1515–1522.
The role of inherited tectono-sedimentary architecture in the develop- Byerlee, J., 1993, Model for episodic flow of high-pressure water in fault
ment of the central Andean mountain belt: Insights from the Cordillera de zones before earthquakes: Geology, v. 21, p. 303–306.
Domeyko: Journal of Structural Geology, v. 30, p. 1520–1539. Caine, J.S, Craig, B., and Forster, C.B., 1996, Fault zone architecture and
Angelier, J., 1994, Fault slip analysis and paleostress reconstruction, in permeability structure: Geology, v. 24, p. 1025–1028.
Hancock, P.L., ed., Continental deformation: Oxford, Pergamon Press, Campbell, I.H., Ballard, J.R., Palin, J.M., Allen, C., and Faunes, A., 2006,
p. 53–100. U-Pb zircon geochronology of granitic rocks from the Chuquicamata-El
Arancibia, O.N., and Clark, A.H., 1996, Early magnetite-amphibole-pla- Abra porphyry copper belt of northern Chile: Excimer laser ablation ICP-
gioclase alteration-mineralization in the Island copper porphyry copper- MS analysis: Economic Geology, v. 101, p. 1327−1344.
gold-molybdenum deposit, British Columbia: Economic Geology, v. 91, Camus, F., 2003, Geología de los Sistemas Porfídicos en los Andes de Chile:
p. 402–438. Santiago, Chile, Servicio Nacional de Geología y Minería, 267 p.
Arnott, A.M., 2003, Evolution of the hydrothermal alteration at the Chuqui- Castillo, P.R., 2012, Adakite petrogenesis: Lithos, v. 134−135, p. 304–316.
camata porphyry copper system, northern Chile: Unpublished Ph.D. thesis, Cathles, L.M., 1977, An analysis of the cooling of intrusives by ground-water
Halifax, Canada, Dalhousie University, 450 p. convection which includes boiling: Economic Geology, v. 72, p. 804–826.
Arnott, A.M., and Zentilli, M., 2003, The Chuquicamata Intrusive Complex: Chen, Y., Wu, J., and Suppe, J., 2019, Southward propagation of Nazca
Its relation to the Fortuna Intrusive Complex, and the role of the Banco subduction along the Andes: Nature, v. 565, p. 441–447, doi: 10.1038/
porphyry in the potassic alteration zone: Actas del X Congreso Geológico s41586-018-0860-1.
Chileno, Concepción, October 6–10, 2003, Proceedings, CD. Cloos, M., and Sapiie, B., 2013, Porphyry copper deposits: Strike-slip faulting
Arriagada, C., Roperch, P., Mpodozis, C., and Cobbold, P.R., 2008, Paleogene and throttling cupolas: International Geology Review, v. 55, p. 43–65.
building of the Bolivian orocline: Tectonic restoration of the central Andes Cobbold, P., Rossello, E.A., Roperch, P., Arriagada, C., Gomez, L.A., and
in 2-D map view: Tectonics, v. 27, doi: 10.1029/2008TC002269. Lima, C., 2007, Distribution, timing, and causes of Andean deformation
Astudillo, N., Roperch, P., Townley, B., and Maksaev, V., 2008, Importance of across South America: Geological Society, London, Special Publication 272,
small-block rotations in damage zones along transcurrent faults: Evidence p. 321–343.
from the Chuquicamata open pit, northern Chile: Tectonophysics, v. 450, Cooke, D.R., Hollings, P., and Walshe, J.L, 2005, Giant porphyry deposits:
p. 1–20. Characteristics, distribution, and tectonic controls: Economic Geology,
Baer, G., Beyth, M., and Reches, Z., 1994, Dikes emplaced into fractured v. 100, p. 801–818.
basement: Journal of Geophysical Research, v. 99, p. 24,039–24,051. Corbett, G., and Leach, T., 1998, Southwest Pacific rim gold-copper systems:
Ballard, J., Palin, J.M., Williams, I., Campbell, I., and Faunes, A., 2001, Two Structure, alteration and mineralization: Society of Economic Geologists,
ages of porphyry intrusion resolved for the super-giant Chuquicamata cop- Special Publication 6, 237 p.
per deposit of northern Chile: Geology, v. 29, p. 383–386. Córdova, S., Demané, E., Fortt, L., Ordenes, T., and Ramírez, F., 2010,
Barra, F., Ruiz, J., Valencia, V.A., Ochoa-Landín, L., Chesley, J.T., and Alteración y mineralización hipógena del yacimiento de Chuquicamata:
Zurcher, L., 2005, Laramide porphyry Cu-Mo mineralization in northern Jornadas de Geociencias, Codelco Norte, Primera Reunión Bi-Anual Geo-
Mexico: Age constraints from Re-Os geochronology in molybdenite: Eco- ciencias 2010, Subgerencia de Geología y Geotecnia, Gerencia de Recursos
nomic Geology, v. 100, p. 1605–1616. Mineros y Desarrollo, Calama, 2010, Proceedings, p. 11–21.
Barra, F., Alcota, H., Rivera, S., Valencia, V., Munizaga, F., and Maksaev, Cornejo, P., Tosdal, R.M., Mpodozis, C., Tomlinson, A.J., Rivera, O., and
V., 2013, Timing and formation of porphyry Cu-Mo mineralization in the Fanning, C.M., 1997, El Salvador, Chile, porphyry copper deposit revisited:
Chuquicamata district, northern Chile: New constraints from the Toki clus- Geologic and geochronologic framework: International Geology Review,
ter: Mineralium Deposita, v. 48, p. 629–651. v. 39, p. 22–54.
Barton, C.A., Zoback, M.D., and Moos, D., 1995, Fluid flow along potentially Couples, G.D., and Lewis, H., 2007, Introduction: The relationship between
active faults in crystalline rock: Geology, v. 12, p. 683–686. damage and localization: Geological Society, London, Special Publication
Beane, R.E., 1974, Biotite stability in the porphyry copper environment: Eco- 289, p. 1–6.
nomic Geology, v. 69, p. 241–256. Cox, S.F., 1995, Faulting processes at high fluid pressures: An example of
Bertrand, G., Guillou-Frottier, L., and Loiselet, C., 2014, Distribution of por- fault-valve behaviour from the Wattle Gully fault, Victoria, Australia: Jour-
phyry copper deposits along the western Tethyan and Andean subduction nal of Geophysical Research, v. 100, p. 841–859.
zones: Insights from a paleotectonic approach: Ore Geology Reviews, v. 60, ——1999, Deformation controls on the dynamics of fluid flow in mesother-
p. 174–190. mal gold systems: Geological Society, London, Special Publication 155,
Birch, F., 1966, Compressibility; elastic constants: Geological Society of p. 123–140.
America Memoirs, v. 97, p. 97–173. ——2005, Coupling between deformation, fluid pressures and fluid flow in
Blanco, N., 2008, Estratigrafía y evolución tectono-sedimentaria de la cuenca ore-producing hydrothermal environments: Economic Geology 100th Anni-
cenozoica de Calama (Chile, 22° S): Unpublished M.Sc. thesis, Spain, Uni- versary Volume, p. 39–75.
versity of Barcelona, 68 p. ——2010, The application of failure mode diagrams for exploring the roles
Blanpied, M.L., Lockner, D.A., and Byerlee, J.D., 1992, An earthquake of fluid pressure and stress states in controlling styles of fracture-controlled
mechanism based on rapid sealing of faults: Nature, v. 358, p. 574–576. permeability enhancement in faults and shear zones: Geofluids, v. 10,
Blundell, R., 2002, The timing and location of major ore deposits in an evolv- p. 217–233.
ing orogen: The geodynamic context: Geological Society, London, Special ——2016, Injection-driven swarm seismicity and permeability enhancement:
Publication 204, p. 1–12. Implications for the dynamics of hydrothermal ore systems in high fluid

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
24 JORGE SKARMETA

flux, overpressured faulting regimes—an invited paper: Economic Geology, Gough, D.I., and Gough, W.I., 1987, Stress near the surface of the Earth:
v. 111, p. 559–587. Annual Review of Earth and Planetary Sciences, v. 15, p. 545–566.
——2020, The dynamics of permeability enhancement and fluid flow in over- Gruen, G., Heinrich, C.A., and Schroeder, K., 2010, The Bingham Can-
pressured, fracture-controlled hydrothermal systems: Reviews in Economic yon porphyry Cu-Mo-Au deposit. II. Vein geometry and ore shell for-
Geology, v. 21, p. 25–82. mation by pressure-driven rock extension: Economic Geology, v. 105,
Cox, S.F., Braun, J., and Knackstedt, M.A., 2001, Principles of structural con- p. 69–90.
trol on permeability and fluid flow in hydrothermal systems: Reviews in Gudmundsson, A., 2011, Rock fractures in geological processes: Cambridge,
Economic Geology, v. 14, p. 1–24. Cambridge University Press, 592 p. 
Cuadra, P., Grez, E., and Gröpper, H., 1997, Geología del yacimiento Guillou-Frottier, L., and Burov, E., 2003, The development and fracturing
Radomiro Tomic: Congreso Geológico Chileno, VIII, Antofagasta, 1997, of plutonic apexes: Implications for porphyry ore deposits: Earth and Plan-
Proceedings, v. 3, p. 1918–1922. etary Science Letters, v. 214, p. 341–356.
Cui, X., Cao, H., Chen, J., Fan, W., and Li, H., 2016, A preliminary study Gustafson, L.B., and Hunt, J.P., 1975, The porphyry copper deposit at El
on the possible range of stress magnitude in the upper part of crust under Salvador, Chile: Economic Geology, v. 70, p. 857–912.
strike-slip faulting regime: International Society for Rock Mechanics and Gustafson, L.B., and Quiroga G.J., 1995, Patterns of mineralization and alter-
Rock Engineering, ISRM-ISRS-2016-050, Finland, May 10–12, 2016, Pro- ation below the porphyry copper orebody at El Salvador, Chile: Economic
ceedings, 8 p. Geology, v. 90, p. 2–16.
De Celles, P.G., Zandt, G., Beck, S.L., Currie, C.A., Ducea, M.N., Kapp, P., Gutscher, M.A., 2002, Andean subduction styles and their effect on thermal
and Schoenbohm, L.M., 2015, Cyclical orogenic processes in the Cenozoic structure and interplate coupling: Journal of South American Earth Sci-
central Andes: Geological Society of America Memoirs, v. 212, p. 459–490. ences, v. 15, p. 3–10.
Deckart, K., Clark, A.H., Aguilar A.C., Vargas R.R., Bertens, N.A., Mortensen, Gutscher, M.-A., Maury, R., Eissen, J.-P., and Bourdon, E., 2000, Can slab
J.K., and Fanning, M., 2005, Magmatic and hydrothermal chronology of melting be caused by flat subduction?: Geology, v. 28, no. 6, p. 535–538.
the giant Río Blanco porphyry copper deposit, central Chile: Implications Harris, A.C., Allen, C.M., Bryan, S.E., Campbell, I.H., Holcombe, R.J., and
of an integrated U-Pb and 40Ar/39Ar database: Economic Geology, v. 100, Palin, J.M., 2004, ELA-ICP-MS U-Pb zircon geochronology of regional
p. 905–934. volcanism hosting the Bajo de la Alumbrera Cu-Au deposit: Implica-
Delaney, P.T., Pollard, D., Zioney, J.I., and McKee., E.H., 1986, Field rela- tions for porphyry-related mineralization: Mineralium Deposita, v. 39,
tions between dykes and joints: Emplacement processes and paleostress p. 46–67.
analysis: Journal of Geophysical Research, v. 91, p. 4920–4938. Hasegawa, A., 2017, Role of H2O in generating subduction zone earthquakes:
Dilles, J.H., and Einaudi, M.T., 1992, Wall-rock alteration and hydrothermal Monographs of Environmental and Earth Planets, v. 5, no. 1, p. 1–34.
flow paths about the Ann-Mason porphyry copper deposit, Nevada; a 6 km Healy, D., 2009, Anisotropy, pore fluid pressure and low angle normal faults:
vertical reconstruction: Economic Geology, v. 87, p. 1963–2001. Journal of Structural Geology, v. 31, p. 561–574.
Dilles, J., Tomlinson, A., Martin, M., and Blanco, N., 1997, El Abra and For- ——2012, Anisotropic poroelasticity and the response of faulted rock to
tuna complexes: A porphyry copper batholith sinistrally displaced by the changes in pore-fluid pressure: Geological Society, London, Special Pub-
Falla Oeste: Congreso Geológico Chileno, VIII, Antofagasta, 1997, Pro- lication 367, p. 201–214.
ceedings, v. 3, p. 1883–1887. Heidrick, T.L., and Titley, S.R., 1982, Fracture and dike patterns in Laramide
Dilles, J.H., Tomlinson, A.J., García, M., and Alcota, H., 2011, The geology of plutons and their structural and tectonic implications: American southwest,
the Fortuna granodiorite complex, Chuquicamata district, northern Chile: in Titley S.R., ed., Advances in geology of the porphyry copper deposits:
Relation to porphyry copper deposits: Society of Geology Applied to Min- Tucson, University of Arizona Press, p. 73–91.
eral Deposits (SGA), Biennial Meeting, 11th, Antofagasta, 2011, Proceed- Heinrich, C.A., 2005, The physical and chemical evolution of low-salinity
ings, p. 399–401. magmatic fluids at the porphyry to epithermal transition: A thermodynamic
Dubé, B., Poulsen, H., and Guha, J., 1989, The effects of layer anisotropy on study: Mineralium Deposita, v. 39, p. 864–889.
auriferous shear zones; the Norbeau mine, Quebec: Economic Geology, Hergert, T., and Heidbach, O., 2011, Geomechanical model of the Marmara
v. 84, p. 871–878. Sea region—II. 3-D contemporary background stress field: Geophysical
Engelder, T., 1987, Joints and shear fractures in rock, in Atkinson, B.K., ed., Journal International, v. 185, no. 3, p. 1090–1102.
Fracture mechanics of rock: London, Academic Press, p. 27–69. Hickman, S., Sibson, R.H., and Bruhn, R., 1995, Introduction to special sec-
——1993, Stress regimes in the lithosphere: Princeton, Princeton University tion: Mechanical involvement of fluids in faulting: Journal of Geophysical
Press, 457 p. Research, v. 100, p. 12,831–12,840.
——1994, Brittle crack propagation, in Hancock, P.L., ed., Continental Hill, D.P., 1977, A model for earthquake swarms: Journal of Geophysical
deformation: Oxford, Pergamon Press, p. 43–52. Research, v. 82, p. 347–352.
England, P., and Molnar, P., 1990, Surface uplift, uplift of rocks and exhuma- Horton, B.K., 2018, Tectonic regimes of the central and southern Andes:
tion of rocks: Geology, v. 3, p. 1173–1177. Responses to variations in plate coupling during subduction: Tectonics,
Etheridge, M.A.,1983, Differential stress magnitudes during regional defor- v. 37, doi: 10.1002/ 2017TC004624.
mation and metamorphism: Upper bound imposed by tensile fracturing: Hubbert, M.K., 1951, Mechanical basis for certain familiar geologic struc-
Geology, v. 11, p. 231–234. tures: Bulletin of the Geological Society of America, v. 62, p. 356–372.
Faulkner, D.R., Mitchell, T.M., Healy, D., and Heap, M.J., 2006, Slip on Jaeger, J.C., 1964, Thermal effects of intrusions: Geophysical Reviews, v. 2,
weak fault by the rotation of regional stress in the fracture damage zone: p. 443–466.
Nature, v. 444, p. 922–925. doi: 10.1038/nature05353. Jaeger, J.C., and Cook, N.G.W., 1979, Fundamentals of rock mechanics, 3rd
Faulkner, D.R., Jackson, C.A.L., Lunn, R.J., Schlische, R.W., Shipton, Z.K., ed.: London, Chapman and Hall, 593 p.
Wibberley, C.A.J., and Withjak, M.O., 2010, A review of recent develop- Jarrell, O.W., 1944, Oxidation at Chuquicamata, Chile: Economic Geology,
ments concerning the structure, mechanics and fluid properties of fault v. 39, p. 251–286.
zones: Journal of Structural Geology, v. 32, p. 1557–1575. Jolly, R.H., and Sanderson, D.J., 1995, Variation in the form and distribution
Faunes, A., Hintze, F., Siña, A., Veliz, H., and Vivanco, M., 2005, Chuquica- of dykes in the Mull swarm, Scotland: Journal of Structural Geology, v. 17,
mata, core of a planetary scale Cu-Mo anomaly, in Porter, T.M., ed., Super p. 1543–1557.
porphyry copper and gold deposits—a global perspective, v. 1: Adelaide, ——1997, A Mohr circle construction for the opening of a pre-existing frac-
PGC Publishing, p. 151–174. ture: Journal of Structural Geology, v. 19, p. 887–892.
Fournier, R.O., 1991, The transition from hydrostatic to greater than hydro- Kahou, Z.S., Brichau, S., Poujol, M., Duchêne, S., Campos, E., Leisen, M.,
static fluid pressure in presently active continental hydrothermal systems in d’Abzac, F., Riquelme, R., and Carretier, S., 2020, First U-Pb LA-ICP-
crystalline rock: Geophysical Research Letters, v. 18, p. 955–958. MS in situ dating of supergene copper mineralization: Case study in the
——1999, Hydrothermal processes related to movement of fluid from plastic Chuquicamata mining district, Atacama Desert, Chile: Mineralium Depos-
into brittle rock in the magmatic-epithermal environment: Economic Geol- ita, doi: 10.1007/s00126-020-00960-2.
ogy, v. 94, p. 1193–1211. Kay, S.M., and Mpodozis, C., 2001, Central Andean ore deposits linked to
Fossen, H., 2016, Structural geology, 2nd ed.: Cambridge, Cambridge Uni- evolving shallow subduction systems and thickening crust: GSA Today,
versity Press, 510 p. March, p. 4–9.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 25

Kay, S.M., Mpodozis, C., and Coira, B., 1999, Magmatism, tectonism, and Mercer, C.N., Reed, M.H., and Mercer, C.M., 2015, Time scales of porphyry
mineral deposits of the central Andes (22°–33°S): Society of Economic Cu deposit formation: Insights from titanium diffusion in quartz: Economic
Geologists, Special Publication 7, p. 27–59. Geology, v. 110, p. 587–602.
Kloppenberg, A., Grocott, J., and Hutchinson, D., 2010, Structural setting Meyer, C., and Hemley, J.J., 1967, Wall rock alteration, in Barnes, H.L.,
and synplutonic fault kinematics of a cordilleran Cu-Au-Mo porphyry min- ed., Geochemistry of hydrothermal ore deposits: New York, Holt,
eralization system, Bingham mining district, Utah: Economic Geology, v. p. 166–23.
105, p. 743–761. Micklethwaite, S., and Cox, S.F., 2004, Fault segment rupture, aftershock
Lamb, S., and Davis, P., 2003, Cenozoic climate change as a possible cause for zone fluid flow, and mineralization: Geology, v. 32, p. 813–816.
the rise of the Andes: Nature, v. 425, p. 792–797. ——2006, Progressive fault triggering and fluid flow in aftershock domains:
Lewis, M., 1996, Characterisation of hypogene covellite assemblages at the Examples from mineralized Archaean fault systems: Earth and Planetary
Chuquicamata porphyry copper deposit, Chile, section 4500N: Unpub- Science Letters, v. 250, p. 318–330.
lished M.Sc. thesis, Halifax, Canada, Dalhousie University, 223 p. Micklethwaite, S., Sheldon, H.S., and Baker, T., 2010, Active fault and shear
Lindsay, D.D., 1998, Structural control and anisotropy of mineraliza- processes and their implications for mineral deposit formation and discov-
tion within the Chuquicamata porphyry copper deposit, northern Chile: ery: Journal of Structural Geology, v. 32, p. 151–165.
Unpublished Ph.D. thesis, Halifax, Canada, Dalhousie University, 371 p. Monecke, T.,  Monecke, J., Reynold, T.J., Tsuruoka, S., Bennett, M.M.,
Lindsay, D.D., Zentilli, M., and Rojas, J., 1995, Evolution of an active duc- Skewes, W.B., and Palin, R.M., 2018, Quartz solubility in the H2O-NaCl
tile to brittle shear system controlling mineralization at the Chuquicamata system: A framework for understanding vein formation in porphyry copper
porphyry copper deposit, northern Chile: International Geology Review, deposits: Economic Geology, v. 113, p. 1007–1046.
v. 37, p. 945–958. Montaño, J.M., 1976, Estudio Geológico de la zona de Caracoles y áreas veci-
Lockner, D.A., 1995, Rock failure: American Geophysical Union Reference nas, con énfasis en el Sistema Jurásico, Provincia de Antofagasta, II Región,
Shelf Series, v. 3, p. 127–147. Chile: Ph.D. thesis, Santiago, Chile, University of Chile, 169 p.
Louks, R.R., 2014, Distinctive composition of copper-ore-forming arc-mag- Moore, D.E., Lockner, D.A., and Byerlee, J.D., 1994, Reduction of perme-
mas: Australian Journal of Earth Sciences, v. 61, p. 5–16. ability in granite at elevated temperatures: Science, v. 265, p. 1558–1561.
Lowell, J.D., and Guilbert, J.M., 1970, Lateral and vertical alteration-min- Morrow, C.A., Moore, D.E., and Lockner, D.A., 2001, Permeability reduc-
eralization zoning in porphyry ore deposits: Economic Geology, v. 65, tion in granite under hydrothermal conditions: Journal of Geophysical
p. 373–408. Research, v. 106, p. 30,551–30,560.
Lowell, R.P., Van Capellan, P., and Germanovich, L.N., 1993, Silica precipita- Mpodozis, C., and Cornejo, P., 2012, Cenozoic tectonics and porphyry cop-
tion in fractures and the evolution of permeability in hydrothermal upflow per systems of the Chilean Andes: Society of Economic Geologists, Special
zones: Science, v. 260, p. 192–194. Publication 16, p. 329–360.
Maksaev, V., 1990, Metallogeny, geological evolution, and thermochronology Mpodozis, C., and Ramos, V.A., 1990, The Andes of Chile and Argentina:
of the Chilean Andes between latitudes 21° and 26° south and the origin Circum-Pacific Council for Energy and Mineral Resources, Earth Science
of mayor porphyry copper deposits: Unpublished Ph.D. thesis, Halifax, Series, v. 11, p. 59–90.
Canada, Dalhousie University, 553 p. Münchmeyer, C., 1996, Exotic deposits—products of lateral migration of
Maksaev, V., and Zentilli, M., 1999, Fission track thermochronology of the supergene solutions from porphyry copper deposits: Society of Economic
Domeyko Cordillera, northern Chile: Implications for Andean tectonics Geologists, Special Publication 5, p. 43–58.
and porphyry copper metallogenesis: Exploration and Mining Geology, v. 8, Munizaga, F., Maksaev, V., Fanning, C.M., Giglio, S., Yaxley, G., and Tas-
p. 65–89. sinari, C.G., 2008, Late Paleozoic-Early Triassic magmatism on the west-
Marinovic, S., and Lahsen, N., 1984, Hoja Calama, Región de Antofagasta: ern margin of Gondwana: Collahuasi area, northern Chile: Gondwana
Santiago, Chile, Servicio Nacional de Geología y Minería, escala 1:250,000. Research, v. 13, p. 407–427.
Masterman, G.J., Cooke, D.R., Berry, R.F., Walshe, J.L., Lee, A.W., and Newhouse, W.H., 1940, Openings due to movement along a curved or irregu-
Clark, A.H., 2005, Fluid chemistry, structural setting, and emplacement lar fault plane: Economic Geology, v. 35, p. 444–464.
Nguyen, P.T., Cox, S.F., Harris, L.B., and Powell, C.M., 1998, Fault-valve
history of the Rosario Cu-Mo porphyry and Cu-Ag-Au epithermal veins,
behaviour in optimally oriented shear zones: An example at the Revenge
Collahuasi district, northern Chile: Economic Geology, v. 100, p. 835–862.
gold mine, Kambalda, Western Australia: Journal of Structural Geology,
May, G., Hartley, A., Chong, G., Stuart, F., and Kape, S., 2005, Eocene to
v. 20, p. 1625–1640.
Pleistocene lithostratigraphy, chronostratigraphy and tecto-sedimentary
Noble, D., McKee, E., and Mégard, F., 1979, Early Tertiary “Incaic” tecto-
evolution of the Calama basin, northern Chile: Revista Geologica de Chile,
nism, uplift, and volcanic activity, Andes of central Peru: Geological Society
v. 32, p. 33–58.
of America Bulletin, v. 90, p. 903–907.
Maydagán, L., Franchini, M., Rusk, B., Lentz, D.R., McFarlane, C., Impic-
Ojala, V.J., Ridley, J.R., Groves, D.I., and Hall, G.C., 1993, The Granny Smith
cini, A., Ríos, F.J., and Rey, R., 2015, Porphyry to epithermal transition
gold deposit: The role of heterogeneous stress distribution at an irregu-
in the Altar Cu-(Au-Mo) deposit, Argentina, studied by cathodolumines-
lar granitoid contact in a greenschist facies terrane: Mineralium Deposita,
cence, LA-ICP-MS, and fluid inclusion analysis: Economic Geology, v. 110, v. 28, p. 409–419.
p. 889–923. Olivares, B., 2001, Alzamiento, termocronometría y evolución tectónica de
McClay, K., Skarmeta, J., and Bertens, A., 2002, Structural controls on por- bloques en la Cordillera de Domeyko, Norte de Chile: Unpublished thesis,
phyry copper deposits in northern Chile: Australian Institute of Geologists, Santiago, University of Chile, 71 p.
Applied Structural Geology for Mineral Exploration and Mining, Kalgoor- Olsen, M.P., Scholz, C.H., and Leger, A., 1998, Healing and sealing of sim-
lie, Western Australia, September 23–25, 2002, Extended Abstracts, p. 127. ulated fault gouge under hydrothermal conditions: Implications for fault
McCuaig, T.C., and Hronsky, J.M., 2014, The mineral system concept: The healing: Journal of Geophysical Research, v. 103, p. 7421–7430.
key to exploration targeting: Society of Economic Geologists, Special Pub- Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P., and Schemman, K.,
lication 18, p. 153–175. 2006, Deformation of the central Andean upper plate system: Facts, fiction
McGarr, A., 1988, On the state of lithospheric stress in the absence of applied and constraints for plateau models, in Oncken, O., Chong, G., Franz, G.,
tectonic forces: Journal of Geophysical Research, v. 93, p. 13,609–13,617. Giese, P., Gotze, H.-J., Ramos, V.A., Strecker, M.R., and Wigger, P., eds.,
——2014, Maximum magnitude earthquakes induced by fluid injection: The Andes: Active subduction orogeny: Frontiers in Earth Sciences, v. 1:
Journal of Geophysical Research, v. 119, p. 1008–1019. Berlin, Springer‐Verlag, p. 3–28.
McGarr, A., and Gay, N.C., 1978, State of stress in the earth’s crust: Annual Ossandon, G., and Zentilli, M., 1997, El distrito de Chuquicamata: Una con-
Review of Earth and Planetary Sciences, v. 6, p. 405–436. centración de cobre de clase mundial: Congreso Geológico Chileno, VIII,
McInnes, B.I.A., Farley, K.A., Sillitoe, R.H., and Kohn, B.P., 1999, Applica- Antofagasta, 1997, Proceedings, v. 3, p. 1888–1892.
tion of apatite (U‐Th)/He thermochronometry to the determination of the Ossandon, G., Freraut, R., Gustafson, L.B., Lindsay, D., and Zentilli, M.,
sense and amount of vertical fault displacement at the Chuquicamata cop- 2001, Geology of the Chuquicamata mine: A progress report: Economic
per deposit, Chile: Economic Geology, v. 94, p. 937–948. Geology, v. 96, p. 249–270.
McKinstry, H.E., 1948, Mining geology: New Jersey, Prentice-Hall, 677 p. Padilla-Garza, R.A., Titley, S.R., and Eastoe, C.J., 2004, Hypogene evolu-
McMillan, W.J., and Panteleyev, A., 1985, Porphyry copper deposits: Geosci- tion of the Escondida porphyry copper deposit, Chile: Society of Economic
ence Canada, Reprint Series, v. 3, p. 45–58. Geologists, Special Publication 11, p. 141–165.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
26 JORGE SKARMETA

Parada, M.A., Aracena, I., and Tanaka, H., 1987, The petrology of Chuquica- Rusk, B.G., Reed, M.H., Dilles, J.H., Klemm, L.M., and Heinrich, C.A., 2004,
mata plutonic complex, Chile: Journal of the Japanese Association of Min- Compositions of magmatic hydrothermal fluids determined by LAICP-MS
eralogy and Petrology, v. 82, p. 177–188. of fluid inclusions from the porphyry copper-molybdenum deposit at Butte,
Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Faral- MT: Chemical Geology, v. 210, p. 173–199.
lon) and South American plates since late Cretaceous time: Tectonics, v. 6, Rusk, B.G., Reed, M.H., and Dilles, J.H., 2008, Fluid inclusion evidence
no. 3, p. 233–248. of magmatic-hydrothermal fluid evolution in the porphyry copper-molyb-
Perry, V.D., 1952, Geology of the Chuquicamata orebody: Mining Engineer- denum deposit at Butte, Montana: Economic Geology, v. 103, no. 2,
ing, v. 4, no. 12, p. 1166–1168. p. 307–334.
Petford, N., 1996, Dykes or diapirs?: Transactions of the Royal Society of Sanderson, D.J., and Zhang, X., 1999, Critical stress localization of flow asso-
Edinburgh: Earth Sciences, v. 87, p. 105–114. ciated with deformation of well-fractured rock masses, with implications
Phillips, W.J., 1972, Hydraulic fracturing and mineralization: Journal of the for mineral deposits: Geological Society, London, Special Publication 155,
Geological Society of London, v. 128, p. 337–359. p. 69–81.
Pilger, R.H., 1984, Cenozoic plate kinematics, subduction and magmatism: Scheuber, E., and Reutter, K.-J., 1992, Magmatic arc tectonics in the central
South American Andes: Journal of the Geological Society of London, v. 141, Andes between 21° and 25° S: Tectonophysics, v. 205, p. 127–140.
p. 793–802. Sdrolias, M., and Muller, R.D., 2006, Controls on back-arc basin formation:
Pinget, M.-C., Dold, B., Zentilli, M., and Fontboté, L., 2015, Reported Geochemistry, Geophysics, Geosystems, v. 7, Q04016, 40 p., doi: 10.1029/
supergene sphalerite rims at the Chuquicamata porphyry deposit (northern 2005GC001090.
Chile) revisited: Evidence for a hypogene origin: Economic Geology, v. 110, Secor, D.T., 1965, Role of fluid pressure in jointing: American Journal of Sci-
p. 253–262. ence, v. 263, p. 633–646.
Proffett, J.M., 2009, High Cu grades in porphyry Cu deposits and their rela- Seedorff, E., Dilles, J.H., Proffett, J.M., Jr., Einaudi, M.T., Zurcher, L.,
tionship to emplacement depth of magmatic sources: Geology, v. 37, no. 8, Stavast, W.J.A., Johnson, D.A., and Barton, M.D., 2005, Porphyry deposits:
p. 675–678. Characteristics and origin of hypogene features: Economic Geology 100th
Rabbia, O.M., Correa, K.J., Hernández, L.B., and Ulrich, T., 2017, “Normal” Anniversary Volume, p. 251–298.
to adakite-like arc magmatism associated with the El Abra porphyry copper Seedorff, E., Barton, M.D., Stavast, W.J., and Maher, D.J., 2008, Root zones
deposit, central Andes, northern Chile: International Journal of Earth Sci- of porphyry systems: Extending the porphyry model to depth: Economic
ences (Geologische Rundschau), doi: 10.1007/s00531-017-1454-0. Geology, v. 103, p. 939–956.
Ramsay, J.G., 1980, The crack-seal mechanism of rock deformation: Nature, Shaw, H.R., 1980, The fracture mechanism of magma transport from the
v. 284, p. 135–139. mantle to the surface, in Hargraves, R.B., ed., Physics of magmatic pro-
Rech, J.A., Currie, B.S., Michalski, G., and Cowan, A, 2006, Neogene cli- cesses: Princeton, Princeton University Press, p. 201–264.
mate change and uplift in the Atacama Desert, Chile: Geology, v. 34, no. 9, Sheldon, H.A., and Micklethwaite, S., 2007, Damage and permeability
p. 761–764. around faults: Implications for mineralization: Geology, v. 34, p. 903–906,
Reed, M.H., 1999, Zoning of metals and early potassic and sericitic hydro- doi: 10.1130/G23860A.
thermal alteration in the Butte, Montana, porphyry Cu Mo deposit: Geo- Sheorey, P., 1994, A theory for in situ stresses in isotropic and trans-
logical Society of America Abstracts with Programs, v. 31, no. 7, p. 381. versely isotropic rock: International Journal of Rock Mechanics, v. 31,
Reutter, K.-J., Scheuber, E., and Helmcke, D., 1991, Structural evidence of p. 23–34.
orogen-parallel strike slip displacements in the Precordillera of northern Sibson, R.H., 1985, A note on fault reactivation: Journal of Structural Geol-
Chile: Geologische Rundschau, v. 80, p. 135–153. ogy, v. 7, no. 6, p. 751–754.
Reutter, K.-J., Scheuber, E., and Chong, G., 1996, The Precordilleran fault ——1990, Rupture nucleation on unfavourably oriented faults: Bulletin of
system of Chuquicamata, northern Chile: Evidence for reversals along arc the Seismological Society of America, v. 80, p. 1580–1604.
parallel strike-slip faults: Tectonophysics, v. 259, p. 213–228. ——1995, Selective fault reactivation during basin inversion: Potential for
Reynolds, P., Ravenhurst, C., Zentilli, M., and Lindsay, D., 1998, High-pre-
fluid redistribution through fault-valve action: Geological Society, London,
cision 40Ar/39Ar dating of two consecutive hydrothermal events in Chuqui-
Special Publication 88, p. 3–19.
camata porphyry cooper system, Chile: Chemical Geology, v. 148, p. 45–60.
——1998, Brittle failure mode plots for compressional and extensional tec-
Richards, J.P., 2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au)
tonic regimes: Journal of Structural Geology, v. 20, p. 655–660.
deposit formation: Economic Geology, v. 98, p. 1515–1533.
——2000, A brittle failure mode plot defining conditions for high-flux flow:
——2009, Post subduction porphyry Cu-Au and epithermal Au deposits:
Economic Geology, v. 95, p. 41–48.
Products of remelting of subduction-modified lithosphere: Geology, v. 37,
——2001, Seismogenic framework for hydrothermal transport and ore depo-
p. 247–250.
sition: Reviews in Economic Geology, v. 14, p. 25–50.
——2011, Magmatic to hydrothermal metal fluxes in convergent and collided
——2004, Controls on maximum fluid overpressure defining condi-
margins: Ore Geology Reviews, v. 40, p. 1–26.
tions for mesozonal mineralisation: Journal of Structural Geology, v. 26,
——2013, Giant ore deposits formed by optimal alignments and combina-
tions of geological processes: Nature Geosciences, v. 6, p. 991–996. p. 1127–1136.
Richards, J.P., Boyce, A.J., and Pringle, M.S., 2001, Geological evolution ——2012, Reverse fault rupturing: Competition between non-optimal and
of the Escondida area, northern Chile: A model for spatial and tempo- optimal fault orientations: Geological Society, London, Special Publication
ral localization of porphyry Cu mineralization: Economic Geology, v. 96, 367, p. 39–50.
p. 271–305. ——2017, The edge of failure: Critical stress overpressure states in different
Ridley, J., 1993, The relations between mean rock stress and fluid flow in the tectonic regimes: Geological Society, London, Special Publication 458, doi:
crust: With reference to vein- and lode-style gold deposits: Ore Geology 10.1144/SP458.5.
Reviews, v. 8, p. 23–37. Sillitoe, R.H., 1973, The tops and bottoms of porphyry copper deposits: Eco-
Rikitake, T., 1959, Studies of the thermal state of the Earth, part 2—heat nomic Geology, v. 68, p. 799–815.
flow associated with magma intrusions: Bulletin of Earthquake Research ——1993, Gold-rich porphyry copper deposits: Geological model and explo-
Institute, Tokyo, v. 37, no. 2, p. 233–243. ration implications: Geological Association of Canada, Special Paper 40,
Rivera, S., Alcota, H., Fontecilla, C., and Kovacic, P., 2009, Supergene modi- p. 465–478.
fication of porphyry columns and the application to exploration with special ——1997, Characteristics and controls of the largest porphyry-cooper gold
reference to the southern part of the Chuquicamata district, Chile: Society and epithermal gold deposits in the circum-Pacific region: Australian Jour-
of Economic Geologists, Special Publication 14, p. 1–14. nal of Earth Sciences, v. 44, p. 373–388.
Rivera, S., Proffett, J., Díaz, J., Leiva, G., and Vergara, M., 2012, Update of ——2010, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.
the geologic setting and porphyry Cu-Mo deposits of the Chuquicamata Sillitoe, R.H., and Hedenquist, J.W., 2003, Linkages between volcanotectonic
district, northern Chile: Society of Economic Geologists, Special Publica- settings, ore-fluid compositions, and epithermal precious metal deposits:
tion 16, p. 19–54. Society of Economic Geologists, Special Publication 10, p. 315–343.
Rusk, B., and Reed, M., 2002, Scanning electron microscope—cathodolumi- Sillitoe, R.H., and McKee, E.H., 1996, Age of supergene oxidation and
nescence analysis of quartz reveals complex growth histories in veins from the enrichment in the Chilean porphyry copper province: Economic Geology,
Butte porphyry copper deposit, Montana: Geology, v. 30, no. 8, p. 727–730. v. 91, p. 164–179.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
STRUCTURAL CONTROLS, CHUQUICAMATA DEPOSIT, CHILE 27

Sillitoe, R.H., and Mortensen, J.K., 2010, Longevity of porphyry copper for- Titley, S.R., Thompson, R.C., Haynes, F.M., Manske, S., Robinson, L.C.,
mation at Quellaveco, Peru: Economic Geology, v. 105, p. 1157–1162. and White, J.L., 1986, Evolution of fractures and alteration in the Sierrita
Sillitoe, R.H., and Perello, J., 2005, Andean copper province: Tectonomag- Esperanza hydrothermal system, Pima County, Arizona: Economic Geol-
matic settings, deposit types, metallogeny, exploration, and discovery: Eco- ogy, v. 81, p. 343–370.
nomic Geology 100th Anniversary Volume, p. 845–890. Tomlinson, A., and Blanco, N., 1997, Structural evolution and displacement
Sillitoe, R.H., Marquardt, J.C., Ramírez, F., Becerra, H., and Gómez, M., history of the West fault system, Precordillera, Chile: Part 2, Post-mineral
1996, Geology of the concealed MM porphyry copper deposit, Chuqui- history: Actas IIX Congreso Geológico Chileno No, Antofagasta, 1997, Pro-
camata district, northern Chile: Society of Economic Geologists, Special ceedings, v. 3, p. 1878–1882.
Publication 5, p. 59–70. ——2008, Geología de la franja El Abra-Chuquicamata, II región (21°45'-
Silver, P.G., Russo, R.M., and Lithgow-Bertelloni, C., 1998, Coupling of 22°30' S): Servicio Nacional de Geología y Minería, Informe Registrado
South American and African plate motions and plate deformation: Science, IR-08-35, 196 p., 5 mapas, escala 1:50,000.
v. 279, p. 60–63. Tomlinson, A.J., Blanco, N., Maksaev, V., Dilles, J.H., Grunder, A.L., and
Simmons, A.T., Tosdal, R.M., Wooden, J.L., Mattos, R., Concha, O., Ladino, M., 2001, Geología de la Precordillera Andina de Quebrada
McCracken, S., and Beale, T., 2013, Punctuated magmatism associated Blanca-Chuquicamata, Regiones I y II (20º30’−22º30’ S): Servicio Nacional
with porphyry Cu-Mo formation in the Paleocene to Eocene of southern de Geología y Minería, Santiago, Informe Registrado IR-01-20, 444 p.
Peru: Economic Geology, v. 108, p. 625–639. Tosdal, R.M., and Dilles, J.H., 2020, Creation of permeability in the porphyry
Singer, D.A., 1995, World class base and precious metal deposits; a quantita- Cu environment: Reviews in Economic Geology, v. 21, p. 173–204.
tive analysis: Economic Geology, v. 90, p. 88–104. Tosdal, R.M., and Richards, J.P., 2001, Magmatic and structural controls on
Singer, D.A., Berger, V.I., and Moring, B.C., 2008, Porphyry copper deposits the development of porphyry Cu ± Mo ± Au deposits: Reviews in Economic
of the world: Database and grade and tonnage models, 2008: U.S. Geo- Geology, v. 14, p. 157–181.
logical Survey, Open-File Report 2008-1155, https://pubs.er.usgs.gov/ Vanderhaeghe, O., and Teyssier, C., 2001, Crustal-scale rheological transi-
publication/ ofr20081155. tions during orogenic collapse: Tectonophysics, v. 335, p. 211–228.
Soh, I., Chang, C., Lee, J., Tae-Hong, L., and Park, E., 2018, Tectonic stress Van Noten, K., Muchez, P., and Sintubin, M., 2011, Stress-state evolution of
orientations and magnitudes, and friction of faults, deduced from earth- the brittle upper crust during compressional tectonic inversion as defined
quake focal mechanism inversions over the Korean Peninsula: Geophysical by successive quartz vein types (High-Ardenne slate belt, Germany): Jour-
Journal International, doi: 10.1093/gji/ggy061. nal of the Geological Society, London, v. 168, p. 407–422.
Somoza, R., 1998, Updated Nazca (Farallon)-South America relative Van Noten, K., Van Balen, H., and Sintubin, M., 2012, The complexity of 3D
motions during the last 40 My: Implications for mountain building in the changes during compresional tectonic inversion at the onset of orogeny:
central Andean region: Journal of South American Earth Sciences, v. 11, Geological Society, London, Special Publication 367, p. 51–69.
p. 211–215. van Wyk de Vries, B., and van Wyk de Vries, M., 2018, Tectonics and vol-
Somoza, R., and Ghidella, M., 2005, Convergencia en el margen occiden- canic and igneous plumbing systems, in Burchardt, S., ed., Volcanic and
tal de América del Sur durante el Cenozoico: Subducción de las placas de igneous plumbing systems: Amsterdam, Elsevier, 356 p., doi: 10.1016/
Nazca, Farallón y Aluk: Revista de la Asociación Geológica Argentina, v. 60, C2015-0-06837-X.
p. 797–809. Vega, M., 1989, Estudio y caracterizacion de vetas de cuarzo en un sector del
——2012, Late Cretaceous to Recent plate motions in western South Amer- yacimiento Chuquicamata: Unpublished B.Sc. thesis, Antofagasta, Chile,
ica revisited: Earth and Planetary Science Letters, v. 331–332, p. 152–163. Universidad de Norte, 77 p.
Somoza, R., Tomlinson, A.J., Zaffarana, S.E., Puigdomenech, C.G., Raposo, Vigneresse, J.-L., Tikoff, B., and Ameglio, L., 1999, Modification of the
M.I.B., and Dilles, J.H., 2015, Tectonic rotations and internal structure of regional stress field by magma intrusion and formation of tabular granitic
Eocene plutons in Chuquicamata, northern Chile: Tectonophysics, v. 654, plutons: Tectonophysics, v. 302, p. 203–224.
p. 113–130. Weinberg, R.F., 1996, Ascent mechanisms of felsic magmas: News and views:
Soto, H., 1979, Alteración y mineralización primaria en Chuquicamata: Transactions of the Royal Society of Edinburgh: Earth Sciences, v. 87,
Unpublished Ph.D. thesis, Spain, Universidad de Salamanca, 233 p. p. 95–103.
Spera, F.J., 1980, Aspects of magma transport, in Hargraves, R.B., ed., Physics Wilkinson, B.H., and Kesler, S.E., 2007, Tectonism and exhumation in con-
of magmatic processes: Princeton, Princeton University Press, p. 265–324. vergent margin orogens: Insights from ore deposits: Journal of Geology,
Stanley, C.R., Holbeck, P.M., Huyck, H.L.O., Lang, J.R., Preto, V.A.G., v. 115, p. 611–627.
Blower, S.J., and Bottaro, J.C., 1995, Geology of the Copper Mountain Wilkinson, W.H., Jr., Vega, L.A., and Titley, S.R., 1982, The geology and ore
alkalic copper-gold porphyry deposits, Princeton, British Columbia: Cana- deposits at Mineral Park, Mohave County, Arizona, in Titley, S.R., ed.,
dian Institute of Mining, Metallurgy and Petroleum, Special Volume 46, Advances in geology of the porphyry copper deposits: Tucson, University of
p. 537–564. Arizona Press, p. 523–541.
Stephens, J.R., Mair, J.L., Oliver, N.H.S., Hart, C.J.R., and Baker, T., Woodcock, N.H., Dickson, D., and Tarasewicz, P.T., 2007, Transient perme-
2004, Structural and mechanical controls on intrusion-related depos- ability and reseal hardening in fault zones: Evidence from dilation breccia
its of the Tombstone gold belt, Yukon, Canada, with comparisons to textures: Geological Society, London, Special Publication 270, p. 43–53.
other vein hosted ore deposit types: Journal of Structural Geology, v. 26, Zentilli, M., Boric, R., Heaman, L., Mathur, R., and Hanley, J., 2015, New
p. 1025–1041. developments on the geology of MMH: Is it the “missing half” of Chuquica-
Taylor, A.V., 1935, Ore deposits at Chuquicamata, Chile: XVI International mata?: XIV Chilean Geological Congress, Symposium, La Serena, October
Geological Congress, Washington, 1935, Proceedings, v. 2, p. 473–484. 4–8, 2015, Extended Abstracts, v. 2, CD.
Tenthorey, E., and Cox, S.F., 2006, Cohesive strengthening of fault zones Zentilli, M., Maksaev, V., Boric, R., and Wilson, J., 2018, Spatial coincidence
during the interseismic period: An experimental study: Journal of Geophys- and similar geochemistry of late Triassic and Eocene-Oligocene magmatism
ical Research, v. 111, B09202, doi: 10.1029/2005JB004122. in the Andes of northern Chile: Evidence from the MMH porphyry type
Tenthorey, E., Cox, S.F., and Todd, H.F, 2003, Evolution of strength recov- Cu-Mo deposit, Chuquicamata district: International Journal of Earth Sci-
ery and permeability during fluid-rock reaction in experimental fault zones: ences, v. 107, p. 1097–1126.
Earth and Planetary Science Letters, v. 206, p. 161–172. Zhang, S., Cox, S.F., and Paterson, M.S., 2001, Microcrack growth and heal-
Terzaghi, R.D., 1965, Sources of errors in joint surveys: Geotechnique, v. 15, ing in deformed calcite aggregates: Tectonophysics, v. 355, p. 17–36.
p. 287–304. Zoback, M., 2010, Reservoir geomechanics: Cambridge, Cambridge Univer-
Thiéblemont, D., Steins, G., and Lescuyer, J.L., 1997, Gisements epither- sity Press, 449 p.
maux et porphyriques: la connexion adakite: Comptes Rendues, Académie Zoback, M.L, and Magee, M., 1991, Stress magnitudes in the crust: Con-
des Sciences de Paris, Earth and Planetary Sciences, v. 325, p. 103–109. straints from stress orientation and relative magnitude data: Philosophi-
Tibaldi, A., 2015, Structure of volcano plumbing systems: A review of multi- cal Transactions, Physical Sciences and Engineering, v. 337, no. 1645,
parametric effects: Journal of Volcanology and Geothermal Research, v. p. 181–194.
298, p. 85–135. Zoback, M.L., Zoback, M.D., Adams, J.,  Assumpção, M., Bell, S.,  et al.,
Titley, S.R.,1993, Characteristic porphyry copper occurrence in the American 1989,  Global patterns of tectonic stress: Nature,  v. 341,  p. 291–298, doi:
Southwest: Geological Association of Canada, Special Paper 40, p. 433–464. 10.1038/341291a0.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user
28 JORGE SKARMETA

Jorge Skarmeta gained a geologist degree at


the University of Chile and Ph.D., M.Sc., and
DIC degrees from the University of London. He
has worked for the Chile Geological Survey, the
National Oil Company (ENAP), and Codelco.
Large parts of his career have been devoted to
structural exploration and mappings in the Andes of
Chile, Peru, and Ecuador, the Cantabrian, Pyrenees, and Rocky Mountains,
the Basin Range (Mexico), and the Brazilian Shield. He has led programs for
the development of 3-D structural models of the large Chilean porphyry cop-
per deposits, development of geographic information and probabilistic area/
target selection systems, and studies of fractured hydrocarbon reservoirs. He
has given talks and courses and currently consults on structural controls of
mineralization and structural models of mineral deposits.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.4769/5192459/4769_skarmeta.pdf


by Carleton University user

You might also like